Você está na página 1de 19

Available online at www.sciencedirect.

com

ScienceDirect
Advances in Space Research 57 (2016) 1585–1603
www.elsevier.com/locate/asr

A parametric sizing model for Molten Regolith Electrolysis reactors


to produce oxygen on the Moon
Samuel S. Schreiner a,⇑,1, Laurent Sibille b,2, Jesus A. Dominguez b,3, Jeffrey A. Hoffman a,4
a
Massachusetts Institute of Technology, 77 Massachusetts Avenue, Cambridge, MA 02139, United States
b
NASA Kennedy Space Center, FL 32899, United States

Received 9 August 2015; received in revised form 4 January 2016; accepted 7 January 2016
Available online 14 January 2016

Abstract

We present a parametric sizing model for a Molten Regolith Electrolysis (MRE) reactor that produces oxygen and molten metals
from lunar regolith. The model has a foundation of regolith material property models validated using data from Apollo samples and
simulants. A multiphysics simulation of an MRE reactor is developed and leveraged to generate a database linking reactor design
and performance trends. A novel design methodology is created which utilizes this database to parametrically design an MRE reactor
that can (1) sustain the required current, operating temperature, and mass of molten regolith to meet a desired oxygen production level,
(2) operate for long periods of time by protecting the reactor walls from the corrosive molten regolith with a layer of solid ‘‘frozen”
regolith, and (3) support a range of electrode separations to enable operational flexibility. Mass, power, and performance estimates
for an MRE reactor are presented for a range of oxygen production levels. Sensitivity analyses are presented for several design variables,
including operating temperature, regolith feedstock composition, and the degree of operational flexibility.
Ó 2016 COSPAR. Published by Elsevier Ltd. All rights reserved.

Keywords: Molten Regolith Electrolysis; In situ resource utilization; Lunar oxygen production; Moon; Molten oxide electrolysis; Parametric modeling

1. Introduction This paradigm has resulted in prohibitive launch costs on


the order of $110,000/kg to the lunar surface (Diaz et al.,
1.1. In-situ resource utilization (ISRU) overview 2005), in 2015 dollars. To enable sustainable, affordable
exploration of the solar system, the reliance on Earth’s
One of the most significant barriers to space exploration resources must be reduced.
is the fact that of the material resources required for a mis- In-situ resource utilization (ISRU) is ‘‘the collection,
sion must be transported from Earth. The rocket equation processing, storing and use of materials encountered in the
(Turner, 2008) describes how a small increase in payload course of human or robotic space exploration that replace
mass results in a dramatic increase launch system mass. materials that would otherwise be brought from Earth”
(Sacksteder and Sanders, 2007). One form of ISRU is pro-
ducing oxygen from lunar regolith, the top layer of loose
⇑ Corresponding author.
material covering the lunar surface. Oxygen is a major
E-mail address: schr0910@umn.edu (S.S. Schreiner).
1
NASA Space Technology Research Fellow, MIT Department of component of launch vehicle, spacecraft, and lander masses
Aeronautics and Astronautics. – 80% of launch vehicle mass is fuel and oxygen, which
2
Surface Systems Group, ESC-5, NASA Kennedy Space Center. itself is 7/8 oxygen by weight (Altenberg, 1990). At the
3
VENCORE-ESC, NASA Kennedy Space Center. same time, oxygen is one of the most abundant lunar
4
Professor of the Practice, MIT Department of Aeronautics and
resources – lunar regolith is more than 40% oxygen by
Astronautics.

http://dx.doi.org/10.1016/j.asr.2016.01.006
0273-1177/Ó 2016 COSPAR. Published by Elsevier Ltd. All rights reserved.
1586 S.S. Schreiner et al. / Advances in Space Research 57 (2016) 1585–1603

weight (Schreiner et al., 2016). The production of this valu- (1963) first suggested using such a process to produce
able resource outside of Earth’s gravity well can support oxygen from lunar soil. Kesterke (1971) conducted the
lunar surface activities and enable orbital refueling to earliest experimental work, using an iridium anode and
reduce the amount of resources that must be launched from silica-carbide cathode to electrolytically produce 170 g of
Earth. oxygen and significant amounts of iron, titanium, alu-
minum, and magnesium from a lunar simulant. In the
1.2. Molten Regolith Electrolysis (MRE) overview ensuing decades, MRE underwent considerable laboratory
development (Lindstrom and Haskin, 1979; Jarrett et al.,
Over twenty different techniques to produce oxygen 1980; Carroll, 1983; Haskin et al., 1992).
from lunar regolith have been proposed in the literature Under the Vision for Space Exploration, further MRE
(Taylor and Carrier, 1993). These techniques range from development continued at MIT (Kim et al., 2011; Wang
solid–gas reactions, such as Hydrogen Reduction of Ilme- et al., 2011), Kennedy Space Center (Sibille et al., 2009),
nite (Gibson and Knudsen, 1985), to molten-gas reactions, Ohio State University (Standish, 2010), and Marshall
such as Carbothermal Reduction (Balasubramaniam et al., Space Flight Center (Curreri et al., 2006). This work
2010), and even direct dissociation via pyrolysis (Senior, demonstrated many suitable anode materials, including
1993). A range of electrochemical techniques have also iridium, platinum, rhodium (Gmitter, 2008), iridium-
been proposed. Some, such Fluxed Magma Electrolysis plated graphite (Paramore, 2010), 50–50 iridium/tungsten
(Keller and Taberaux, 1991), require fluxing reagents to alloys (Vai et al., 2010), and iron–chromium alloys
be mixed with molten regolith to support the electrochem- (Allanore et al., 2013). Reactor designs were scaled up from
ical process. The FFC–Cambridge process (Chen et al., 0.1 A with 0.3 cm2 electrodes to 10 A with 10 cm2 elec-
2000; Schwandt et al., 2012) directly reduces the solid oxi- trodes (Vai et al., 2010; Sirk et al., 2010), making progress
des in lunar regolith using a supporting molten salt. from laboratory-scale testing towards technology demon-
A promising electrochemical technique, called Molten stration levels.
Regolith Electrolysis (MRE) requires no additives, but
rather performs direct electrolysis on raw molten lunar
1.4. Molten regolith containment: joule heated cold wall
regolith (Kesterke, 1971). In the MRE process (also known
operation
as Molten Oxide Electrolysis), granular lunar regolith is fed
into a reactor and heated to a molten state. Molten lunar
Perhaps the most challenging design consideration in an
regolith is conductive enough to sustain electrolysis, in
MRE reactor is the containment of corrosive molten rego-
which a voltage is applied across two electrodes immersed
lith. The longest laboratory experiments lasted on the order
in the molten regolith to drive a current. Oxygen gas is pro-
of a few hours (Paramore, 2010) before molten regolith
duced at the anode and molten iron, silicon, aluminum,
eroded through the inner crucible. Paramore (2010) noted
and titanium, are produced at the cathode.
that ‘‘fortunately, this frustration [crucible failure] is purely
There is a strong impetus to explore the feasibility of an
an artifact of laboratory-scale experimentation”.
ISRU system based on the MRE process. A study compar-
To solve the challenge of molten regolith containment,
ing oxygen production methods, including water extraction
an MRE reactor can be designed similarly to industrial
from lunar polar craters, identified MRE as having the
Hall–Heroult reactors to support joule heated cold wall
most favorable power consumption and the second most
operation (Carroll, 1983; Sibille and Dominguez, 2012).
favorable mass throughput (Mason, 1992). MRE does
In joule heated cold wall operation, the current flowing
not require additional process materials (such as hydrogen,
through the molten regolith generates joule heating (I 2 R)
methane, or fluxing reagents) or the systems required to
to maintain a central molten core, and the thermal gradi-
recycle them, leading to a decrease in system mass and
ents are designed such that the reactor walls are protected
complexity. MRE can extract up to 95% of the oxygen
from the molten core by a layer of solid ‘‘frozen” regolith
from lunar regolith, which decreases regolith throughput
to enable long duration operation as shown in Fig. 1.
requirements and reactor mass and power. With post-
The current streamlines are depicted in red and simplified
reactor processing, the byproducts of MRE (molten iron,
anodic and cathodic reactions are indicated. Oxygen is pro-
silicon, aluminum, titanium, and glassy slag) can be used
duced at the anode and exits the reactor through a port
to produce infrastructure, spare parts and even solar arrays
(not shown).
on the lunar surface (Landis, 2005; Curreri et al., 2006). In
fact, MRE is also being developed for environmentally-
friendly metal production on Earth with a zero-carbon 1.5. Motivation for MRE reactor modeling
footprint (Gmitter, 2008; Vai et al., 2010).
Although a number of analyses have examined the util-
1.3. Historical development of MRE ity of lunar oxygen production (Simon, 1985; Joosten and
Guerra, 1993; Duke et al., 2003), the accuracy of their find-
The process of electrolytically extracting metals from ings depends heavily on the mass and performance of the
molten ores was first patented by Aiken (1906) and Carr ISRU systems involved. Sherwood and Woodcock (1993)
S.S. Schreiner et al. / Advances in Space Research 57 (2016) 1585–1603 1587

terrestrial reactors, which results in significantly different


reactor behavior. The differences in process chemistry
between lunar and terrestrial techniques significantly affects
reactor design (Inman and White, 1978), inhibiting the
ability to translate models for terrestrial reactors to lunar
applications (Poizeau and Sadoway, 2011). Furthermore,
in lunar applications the thermal behavior of the external
reactor surface will consist only of radiative heat transfer
and may include multi-layer insulation, which will decrease
surface emissivity as compared to terrestrial reactors.
There has been some preliminary simulation of lunar
MRE reactors in the literature. Sibille and Dominguez
(2012) demonstrated that a joule heated cold wall MRE
Fig. 1. Schematic of a Molten Regolith Electrolysis reactor that produces
reactor appears feasible with lunar regolith feedstock, but
oxygen gas at the anode and molten metals at the cathode. The central they studied a point design rather than developing a para-
molten core is insulated by solid ‘‘frozen” regolith around the reactor metric model capable of studying reactor performance
perimeter. across a range of production levels. Previous work (both
simulation and experimental) has studied reactors that
conducted an extensive economic analysis and calculated a use tens of amperes of current, while production-level reac-
cost of $18,370/kg (2015 dollars) to produce oxygen on the tors will need to use on the order of kiloamperes. The
lunar surface using the Hydrogen Reduction of Ilmenite proper method for scaling up MRE reactor designs to these
process. However, they noted that, ‘‘the sensitivities [of higher production levels remains to be explored.
their economic model] are modest, except for the mass of
production hardware”. When the ISRU system mass was 1.6. MRE reactor model overview
varied by a factor of two, the oxygen production cost
varied from $12,570/kg to $29,857/kg (2015 dollars). To this end, we develop a sizing model for an MRE
Simon (1985) showed that, after the Earth-to-Moon trans- reactor that parametrically generates reactor design and
portation cost, the power required for lunar ISRU had the performance estimates for a set of inputs, including oxygen
biggest impact on its economic feasibility. To understand production level, operating temperature, and regolith feed-
the appropriate applications of lunar ISRU, it is imperative stock type. This model can be used to (1) quantitatively
to accurately model the mass, power and performance of compare MRE to other lunar ISRU techniques, (2) study
such systems. how MRE reactors scale to higher production levels, and
Furthermore, there is a large impetus to model MRE (3) improve our understanding of optimal process condi-
reactors. Altenberg (1990) noted that two of the primary tions and feedstock requirements.
issues associated with MRE were (1) the lack of specified The MRE reactor model has a foundation of lunar rego-
optimal process conditions, feed rate, and feedstock lith material property models which were validated using
requirements and (2) a poor understanding of the meaning- data from Apollo samples and regolith simulants
ful design parameters and oxygen extraction efficiency. (Schreiner et al., 2016). The reactor model calculates the
Teeple (1994) surmised that ‘‘the electrolysis techniques [in- amount of regolith (Section 2) and current (Section 3)
cluding MRE], involve high temperatures, so one would required to meet the desired oxygen production level. A
expect high plant masses”. Uncertainty on the operating multiphysics simulation, which models the electrochemical
conditions and mass and power of an MRE reactor can and thermodynamic behavior of an MRE reactor, is devel-
be reduced through modeling. oped and leveraged to create an extensive tradespace of
Although parametric models to predict the mass reactor designs (Section 4).
and power of reactors utilizing other lunar ISRU processes As described in Section 5, the reactor model parametri-
have been developed (Hegde et al., 2009; Balasubramaniam cally generates a reactor design that (1) sustains the
et al., 2010), a similar model of suitable fidelity for a lunar required current, operating temperature, and mass of mol-
MRE reactor does not yet exist in the literature. This deficit ten regolith to meet a desired oxygen production level, (2)
has prevented quantitative comparisons between MRE and ensures that the reactor walls are insulated from the molten
other processing techniques (Chepko, 2009). A substantial core by a layer of solid lunar regolith during joule heated
amount of work modeling electrochemical reactors exists in cold wall operation (see Section 1.4), and (3) enables a
the literature (Inman and White, 1978; Gerogiorgis and range of viable electrode separations to support flexible
Ydstie, 2005; Kar and Evans, 2008; Kennedy, 2012), but reactor operation. The methodology for estimating reactor
this work has primarily focused on terrestrial reactors power and mass is presented in Section 6. The mass, power,
which differ from lunar reactors in a number of ways and performance estimates are given for a range of oxygen
(Carroll, 1983). MRE reactors operate on raw molten production levels in Section 7, including sensitivity analyses
regolith without the fluxing reagents commonly used in for several design variables.
1588 S.S. Schreiner et al. / Advances in Space Research 57 (2016) 1585–1603

2. Calculating regolith processing requirements tion of polymer chains to produce oxygen in the neutral
state, but Eq. (1) still represents the correct stoichiometry.
Lunar regolith is primarily comprised of plagioclase, The MRE model calculates the regolith processing
pyroxene, olivine, and ilmenite, and each of these minerals requirement according to the reactions shown in Eq. (1)
are composed of oxides, including iron(II) oxide (FeO), sil- adapted for each oxide specie in lunar regolith. The
ica (SiO2), alumina (Al2O3), titania (TiO2), magnesia amount of oxygen extracted per kilogram regolith, termed
(MgO), and calcium oxide (CaO). Data from Schreiner the ‘‘oxygen extraction yield” (kO2 ), is calculated as:
et al. (2016) for the oxide compositions of three different X  
mO2 MW O2
types of lunar regolith were used to characterize the rego- k O2 , ¼ ðwi Þ ðrmol;i Þðefrac;i Þ; ð2Þ
lith feedstock in the MRE reactor model. Fig. 2 shows mregolith i
MW oxide;i
the composition data from Schreiner et al. (2016) with a
map of the lunar surface with the lunar High-Ti Mare (yel- where mO2 is the mass of oxygen produced, mregolith is the
low), Low-Ti Mare (cyan) and Highlands (older forma- mass of regolith processed, wi is the weight percent of oxide
tions in red, younger formations in blue), from the i in lunar regolith from Schreiner et al. (2016), MW oxide;i is
Clementine UVVIS instrument (Lucey et al., 2000). the molecular weight of oxide i; rmol;i is the number of moles
The MRE process utilizes a current flowing through of oxygen in one mole of oxide i, and efrac;i is the fraction of
molten regolith to electrochemically produce oxygen. The oxide i that is reduced in each batch (between 0 and 1, to
reactions at the interface between the electrodes and the account for the fact that certain oxides may be partially
raw molten regolith are presented below in simplified form reduced or not reduced at all). In the MRE reactor model,
using the reduction of iron oxide (FeO) as an example Eq. (2) is used to calculate the amount of regolith that must
(Kesterke, 1971; Colson and Haskin, 1990): be processed to meet a given oxygen demand.
The fraction of each specie that can be electrolyzed
½Net Cathode Reaction Fe2þ þ 2e ! Fe0 (efrac;i ) is dependent upon the operating temperature,
ð1Þ because the liquidus temperature of molten regolith gener-
½Net Anode Reaction 2O2 ! O2 þ 4e
ally increases as the composition changes during electroly-
sis. Fig. 3 shows the liquidus temperature throughout the
Colson and Haskin (1990) noted that, due to the slow batch electrolysis calculated using phase diagrams from
kinetics of O2 formation from silicate polymer chains, Slag Atlas (Allibert et al., 1995). To capture the time-
the actual anodic reaction will primarily involve the oxida- varying behavior throughout a batch, the model assumes

Fig. 2. The composition of three different types of lunar regolith: High-Titanium Mare (yellow), Low-Titanium Mare (cyan), and Highlands (older rock in
red, younger rock in blue), from Schreiner et al. (2016). Composition data from Apollo and Luna missions (Stoeser et al., 2010) and imagery data from
Clementine UVVIS instrument (Lucey et al., 2000). (For interpretation of the references to color in this figure legend, the reader is referred to the web
version of this article.)
S.S. Schreiner et al. / Advances in Space Research 57 (2016) 1585–1603 1589

Fig. 3. As electrolysis progresses, the composition and properties of the molten regolith vary. (top) The liquidus temperature, calculated from Slag Atlas
(Allibert et al., 1995), for Low-Ti Mare lunar regolith initially increases as electrolysis progresses. As the operating temperature increases, the reactor can
extract more oxygen per unit regolith. (bottom) Highlands regolith initially has a higher liquidus temperature than Mare, but as electrolysis progresses the
liquidus temperature for Mare regolith rises dramatically higher.

that oxide species are discretely reduced in the order of MgO and TiO2 leads to a general decrease in liquidus tem-
increasingly negative Gibbs Free Energy. As depicted in perature. Finally, as Al2O3 is reduced the liquidus temper-
the top plot, after FeO is reduced, the liquidus temperature ature first decreases and then increases to a plateau for a
increases as SiO2 is reduced. After SiO2, the reduction of mixture dominated by the remaining CaO.
1590 S.S. Schreiner et al. / Advances in Space Research 57 (2016) 1585–1603

The bottom plot in Fig. 3 compares the liquidus temper- where nO2 is the desired moles of diatomic oxygen, n is the
ature curves for Mare and Highlands lunar regolith, which number of electrons required to reduce diatomic oxygen
vary due to differences in composition. Although the High- (=4), F is Faraday’s constant, and gI is the average current
lands regolith has a higher liquidus temperature early in the efficiency defined as the ratio of the oxygen mass produced
batch, the liquidus temperature for Mare regolith quickly to the theoretical prediction of Faraday’s law.
rises above that of Highlands. This is due to the fact that Using Eq. (2) to substitute for mO2 and including the
Mare regolith has more MgO while Highlands regolith oxide current efficiencies in the summation yields:
has more Al2O3 (see Fig. 2). X ðwi Þðrmol;i Þðefrac;i Þ
Electrolysis is allowed to progress until the liquidus C ¼ ðmregolith ÞðnF Þ ; ð5Þ
temperature comes within a 50 K safety margin of the i
ðMW oxide;i ÞðgI;i Þ
operating temperature. At this point fresh regolith is added
where gI;i is the expected current efficiency while oxide i is
while the leftover slag and molten metals are removed, and
the primary specie being reduced.
the batch process begins again. As shown in the top plot of
From experimental work conducted at MIT, the
Fig. 3, with an operating temperature of 1900 K, the reac-
expected current efficiency while electrolyzing iron-
tor can reduce approximately 1/2 of the SiO2 for High-Ti
bearing molten mixtures is 30–60% (Sirk et al., 2010) due
Mare regolith (kO2 ¼ 0:19 kg O2/kg regolith) and High-
to the redox cycling of multivalent ions (Yen, 1977; Chen
lands regolith (kO2 ¼ 0:13) before needing to start a new
et al., 2000). As a conservative estimate, the lower bound
batch with fresh regolith. If the operating temperature is
of 30% is used for FeO. A current efficiency of 50% is used
raised to 2100 K, the reactor can now reduce 5/6 of the
for TiO2 due to a similar redox cycling phenomena (Chen
SiO2 for High-Ti Mare (kO2 ¼ 0:24), and all of the SiO2,
et al., 2000). Na2O, P2O5, K2O, and MgO may form gas-
MgO, and TiO2 for Highlands (kO2 ¼ 0:35) Thus, higher
eous products when reduced (Lindstrom and Haskin,
operating temperatures increase the oxygen extraction yield
1979; Sirk et al., 2010), which will likely recombine with
and enable a more diverse set of oxides to be reduced. For
some of the product O2 gas above the melt. This cyclic
all species, efrac;i had a maximum of 0.95 to account for the
behavior is captured with an estimated current efficiency
fact that when the concentration of a specie becomes too of 50%, though this effect can be mitigated with an oxygen
low, the specie will no longer be preferentially reduced in collection tube near the anode (Sirk et al., 2010). For all
favor of more abundant species. other species in lunar regolith, a current efficiency of close
In addition to avoiding solidification of the molten core, to 100% can be expected (Sibille et al., 2010) but a value of
there is another important aspect in the batch operation of 95% is used as a conservative estimate.
an MRE reactor. As FeO and SiO2 are reduced, the residual Using the number of coulombs predicted by Eq. (5), the
melt moves towards the spinel stability field and a MgAl2O4 average current (I avg ) over an entire batch can be
spinel begins to precipitate (Colson and Haskin, 1990). This
calculated:
spinel must be periodically removed from the reactor to
enable further regolith processing. Future work can investi- C
I avg ¼ ; ð6Þ
gate methods for managing spinel production, either ðtbatch  tfeed Þ
through mechanical or electrochemical means.
Using the fraction electrolyzed (efrac;i ) of each oxide in where tbatch is the total batch time and tfeed is an assumed
lunar regolith, the amount of molten metal produced can 5 min downtime between each batch during which electrol-
also be calculated: ysis is not performed, to be conservative.
  The model assumes that the reactor will be operated in a
MW metal;i constant power mode to minimize peak power. Constant
mmetal;i ¼ ðmregolith Þðwi Þ ðefrac;i Þ; ð3Þ
MW oxide;i power is achieved by varying the current (about an average
value of I avg ) as the voltage changes due to different species
where mmetal;i is the mass of metal produced from oxide i
being electrolyzed and the conductivity of the melt chang-
and MW metal;i is the molecular weight of the product metal
ing throughout the batch. For instance, for an oxygen pro-
i.
duction level of 500 kg/year, the model predicts a required
operating voltage of around 0.96 V while FeO is the pri-
3. Calculating reactor current
mary object of electrolysis and 5.61 V while SiO2 is the pri-
mary object of electrolysis. The respective operating
The MRE reactor model also calculates the current
current for these two species is 1470 A and 250 A. Both
required to meet the desired oxygen production level. Fara-
of these operating conditions result in a power consump-
day’s law (Faraday, 1834), with an adjustment for the
tion of 1.4 kW.
expected current efficiency, is used to calculate the number
of coulombs (C) required to produce a given mass of
oxygen: 4. Multiphysics reactor simulation
    
nF mO2 nF A multiphysics simulation, presented in Schreiner et al.
C ¼ ðnO2 Þ ¼ ; ð4Þ
gI MW O2 gI (2015), was developed to predict the coupled electrochemical
S.S. Schreiner et al. / Advances in Space Research 57 (2016) 1585–1603 1591

and thermodynamic behavior of an MRE reactor. Fig. 4 As described in Section 3, to maintain constant power the
shows a cross-sectional side view of a simulated reactor, current can be varied inversely with voltage (Ptotal = I  V
which is revolved around the dashed axis on the left to gen- is constant). Heat is generated in the molten region by resis-
erate a 3D reactor design (shown in Fig. 1). The rainbow line tive heating from the over-voltage above the decomposition
illustrates the phase transition between the inner core of mol- potential (Vover) multiplied by the current (Pheat = -
ten regolith and the outer solid ‘‘frozen” regolith layer that I  Vover). Constant power operation does not ensure that
insulates the reactor walls. Current streamlines in the molten the heating power in the reactor is constant over time, but
regolith and electrodes are depicted in black. it does reduce the variation in this value. Even as heating
Oxygen is produced at the anode, travels through the power is held relatively constant, the endothermic heatsink
overhead cavity, and leaves the reactor through an exit will vary over time as different oxide species are reduced.
port. Due to the fact that this study was limited to an Thus, the simulation presented in this section represents a
axis-symmetric simulation, this exit port was not included. steady-state approximation of the time-varying behavior
Furthermore, modeling oxygen gas flow in the simulation of the reactor. Future work will need to study the complex,
dramatically impeded convergence and was thus omitted time-varying behavior of an MRE reactor.
to enable the large parametric sweeps described below.
There are four primary modes of heat transfer in the 5. Using the simulation to guide reactor design
simulation: conduction (the blue arrows in Fig. 4),
surface-to-surface radiation in the overhead cavity There are four primary criteria that must be satisfied by
(magenta arrows), radiation to ambient on the exterior sur- an MRE reactor design. The first two, the required molten
face (black arrows), and radiation in participating media mass and current in the reactor, are derived from the oxy-
(red arrows). Radiation in participating media models the gen production level using Eqs. (2) and (6), respectively.
emission, absorption, and scattering of radiation within The third, operating temperature, is a user input. The
the high temperature regolith (Arndt et al., 1979; Sibille fourth criteria is that the reactor design can sustain the
and Dominguez, 2012). joule heated cold wall operation described in Section 1.4.
As presented in Schreiner et al. (2015), data on the To generate a reactor design that meets these performance
Enthalpy of Formation, Gibbs Free Energy, specific heat criteria, the reactor diameter, wall thermal conductivity,
and latent heat of melting for lunar regolith summarized and electrode separation must be carefully designed. This
in Schreiner et al. (2016) were leveraged to add heatsinks section gives an overview of a novel design methodology
for the endothermic chemical electrolysis and the heating developed to address this challenge.
of newly added regolith in molten region.
The multiphysics simulation is a steady-state approxima- 5.1. Effects of molten mass and operating temperature: the
tion of the time-varying behavior of the reactor. Although cutoff line
the thermal performance of the reactor will vary over time
as different oxide species are reduced, this variance is some- The multiphysics simulation described in Section 4 was
what mitigated by the constant power operational concept. leveraged to generate a database of over 100,000 reactor

Fig. 4. A cross-section view of the cylindrical MRE reactor in the multiphysics simulation. The heat fluxes modeled are shown with different colored
arrows, scaled logarithmically by the same factor. The 3D reactor geometry is generated by revolving this cross section around the axis of symmetry on the
left. (For interpretation of the references to colour in this figure legend, the reader is referred to the web version of this article.)
1592 S.S. Schreiner et al. / Advances in Space Research 57 (2016) 1585–1603

designs (consisting of unique combinations of reactor It is critical to note that the cutoff lines illustrated in
diameter, electrode separation, wall thermal conductivity, Fig. 5 are for a fixed current and wall thermal conductivity.
current, and regolith feedstock type) with associated per- The range of designs simulated in the parametric sweep
formance characteristics (molten mass within the reactor, described in Section 4 included reactor currents from
operating temperature, and whether or not the joule heated 250 A to 3 kA and wall conductivities from 0.05 W/m-K
cold wall condition was satisfied). A small sample of the to 5 W/m-K. It was observed that current and wall conduc-
data for a representative reactor design, with 2 kA of cur- tivity also affect reactor performance and the cutoff lines,
rent and a wall thermal conductivity of 5.0 W/m-K, is but for the sake of conciseness, these plots are not included
shown in Fig. 5. The mass of molten regolith in the reactor here. The reader is directed to Schreiner (2015) for a more
(left) and operating temperature (right) are plotted over a detailed discussion of the effects of current and wall ther-
range of reactor diameters, where each line represents a dif- mal conductivity on reactor performance.
ferent electrode separation. The molten mass on the left of
Fig. 5 first rises with the diameter squared because all of the 5.2. Designing the electrode separation
regolith within the reactor is molten. Beyond a certain
inflection point, as reactor diameter increases solid regolith As is evident in the left-hand plot in Fig. 5, as the reac-
begins forming at the outer wall of the reactor, decreasing tor diameter increases the electrode separation must also
the mass of molten regolith in the reactor as discussed by increase (moving from the green line towards the black
Schreiner (2015). one) in order to stay on the cutoff line and maintain joule
Data points in Fig. 5 with a red ‘‘X” overlaid represent heated cold wall operation. The database generated using
infeasible designs, in which the reactor wall temperature the reactor simulation was fit with an empirical equation
rises above 1400 K and becomes too close to the melting for the electrode separation that enables joule heated cold
temperature of lunar regolith circa 1500 K. Lower cutoff wall operation for a given reactor design of diameter (D),
temperatures were evaluated by Schreiner (2015) and found current (I) and wall thermal conductivity (k wall ):
to be less optimal in terms of both reactor mass and power. 0 1
The reactor design is chosen using the line that separates
B e3 ðD  e2 Þ C
De ¼ e1 I e8 ðk wall Þ 7 exp @ 
e
infeasible and feasible design regions, termed the ‘‘cutoff e 6 A; ð7Þ
line” as shown in Fig. 5. The choice to design reactors on
e4
I e5 kwall þ1
the cutoff lines was made after an initial analysis revealed
that those designs had the lowest heat loss per kilogram where ei are regression coefficients (shown in Table 1)
of molten regolith within the reactor (Schreiner, 2015). which depend upon regolith type.

Fig. 5. The molten mass in an MRE reactor (left) and the operating temperature (right) depend on reactor diameter and electrode separation. Red X’s
indicate infeasible designs for which the side wall temperature comes within 100 K of the melting temperature of regolith (1500 K). The ‘‘cutoff line”
separates the infeasible and feasible design regions. (For interpretation of the references to color in this figure legend, the reader is referred to the web
version of this article.)
S.S. Schreiner et al. / Advances in Space Research 57 (2016) 1585–1603 1593

Table 1
The regression coefficients for Eq. (7), which predicts the appropriate
electrode separation needed to maintain thermal equilibrium in an MRE
reactor.
Regolith e1 e2 e3 e4
Highlands 0.0011 0.1945 2.5738 0.6487
Mare 0.0008 0.0458 4.6638 1.3147
Regolith e5 e6 e7 e8
Highlands 0.4656 0.3522 0.0570 0.4348
Mare 0.5827 0.3243 0.0064 0.5857

The electrode separation scaling equation (Eq. (7)) is


plotted in Fig. 6 against a small sample of the data from
the multiphysics simulation. The required electrode separa-
tion increases with reactor diameter, decreases with current
(left plot), and increases with wall conductivity (right plot)
to ensure that the reactor does not generate too much heat
and violate the joule heated cold wall constraint described
in Section 1.4.

5.3. Bounding reactor diameter and electrode separation

The cutoff lines in Fig. 5 prove useful for defining bounds


on the reactor design, as shown in Fig. 7. Succinctly, (1) the
molten mass requirement sets a minimum feasible reactor
diameter, Dmin , (2) the operating temperature sets a maxi-
mum feasible reactor diameter, Dmax , and (3) Dmax and
Dmin depend on the reactor current and the wall thermal
conductivity. The cutoff line represents the minimal reactor
diameter that can meet the required amount of molten rego-
lith in the reactor while ensuring that the molten regolith
core is insulated from the reactor walls by a layer of
‘‘frozen” solid regolith (left plot in Fig. 7). As shown in
Fig. 5, decreasing the reactor diameter (to the left of the cut-
off line) crosses into the infeasible region.
As can be seen in the right plot Fig. 7, the cutoff line rep- Fig. 6. The required electrode separation for a given reactor design. The
resents the maximum reactor diameter that can meet the electrode separation is affected by current (top) and wall thermal
required operating temperature in the reactor while ensuring conductivity (bottom). A regression model, illustrated by the solid line,
was fit to the multiphysics data, depicted by the data points connected by
that no molten regolith touches the reactor wall. As shown in
dashed lines (R2=0.975, RMSE = 0.0032 m).
the right plot of Fig. 5, increasing the diameter to the right of
the cutoff line crosses into the infeasible region, meaning that
molten regolith will touch the reactor wall.
Fig. 7 depicts a scenario in which 40 kg of molten mass in The molten mass and operating temperature data from
the reactor and a current of 2.5 kA are required to meet the the multiphysics simulation (of which only a small portion
desired oxygen production level and the user sets an operat- is shown in Figs. 5–7) were fit with empirical regression
ing temperature of 2250 K. The 40 kg of molten mass equations to predict the molten mass (M m ) and operating
imposes a minimum reactor diameter of 1.27 m for a current temperature (T op ) for a given reactor design:
of 2.5 kA (left plot in Fig. 7). The 2250 K operating temper- 0 1
m ð D  m Þ
ature imposes a maximum feasible diameter of 1.4 m for a M m ¼ I m6 m4 exp @  A;
1 5
ð8Þ
current of 2.5 kA (right plot in Fig. 7). As described in Sec- I m3
m2 þ kwall þ0:15
1
tion 5.1, the cutoff lines are a function of wall thermal con- 0 1
ductivity, so the bounds on reactor diameter can be t ð D  t Þ
T op ¼ t1 þ I t5 t2 exp @  A;
3 7
manipulated by varying that design parameter. Due to the ð9Þ
fact that there is a one-to-one, monotonic mapping between I t6 1 þ kwalltþ0:154

diameter and electrode separation, the bounds on reactor


where mi and ti are regression coefficients which depend
diameter also translate into bounds on the electrode separa-
upon the type of regolith as shown in Table 2.
tion, calculated by substituting Dmin and Dmax into Eq. (7).
1594 S.S. Schreiner et al. / Advances in Space Research 57 (2016) 1585–1603

Fig. 7. An example of how the MRE model calculates the bounds on reactor diameter. In this example (1) for a given reactor current of 2.5 kA, (2) the
required molten mass in the reactor (40 kg) sets a lower bound on the reactor diameter of 1.27 m (left plot) and (3) the desired operating temperature
(2250 K) imposes an upper bound on reactor diameter of 1.4 m (right plot). This example is for a reactor operating on High-Ti Mare regolith with a wall
thermal conductivity of 5.0 W/m-K.

Eqs. (8) and (9) can be algebraically manipulated to Due to the fact that the operating temperature (T op ) is
derive expressions for the minimum and maximum diame- set by the user, and the molten mass (M m ) and current (I)
ter bounds illustrated in Fig. 7: are fixed for a given oxygen production level and batch
   time (via Eqs. (2) and (6), respectively), the thermal con-
I m3 Mm 1
Dmin ¼ m5 þ log m2 þ ð10Þ ductivity of the wall is the only free parameter in Eq.
m1 m4 I m6 k wall þ 0:15
   (12). By substituting Eqs. (10) and (11) into Eq. (12), a
I t6 T op  t1 t4 closed-form expression for the required wall thermal con-
Dmax ¼ t7  log 1 þ ð11Þ
t3 I t5 t 2 k wall þ 0:15 ductivity can be derived:
 m  
T t
I m3 Ut3 log mM4 Im  I t6 m1 t4 log Iopt5 t2 1
6

5.4. Enabling operational flexibility: the design margin k wall ¼ ;


b
 
During reactor operation, the electrode separation must T op  t1
where b  I t6 m1 log  m1 t3 t7 . . . þ Um1 m5 t3
be varied to control the thermal behavior of the reactor. To I t5 t 2
 
quantify operational flexibility, a new design variable was Mm
introduced, termed the ‘‘design margin” (U): þ I m3 Um2 t3 log  0:15 ð13Þ
m4 I m6
Dmax ðT op ; I; k wall Þ The thermal conductivity prescribed by Eq. (13), with the
U, ; ð12Þ
Dmin ðMM; I; k wall Þ maximum diameter from Eq. (11), (1) satisfies the molten
which represents the ratio of the maximum acceptable mass, current and operating temperature constraints, (2)
diameter from Eq. (11) to the minimum acceptable sets the maximum and minimum diameter bounds within
diameter from Eq. (10). Imposing the constraint that a certain range of one another as specified by the design
U P 1 ensures that the diameter bounds are properly margin, and (3) ensures that the reactor can support joule
ordered and U ¼ 1 results in only a single acceptable value heated cold wall operation (no molten regolith touches the
for reactor diameter. reactor wall).

Table 2
The regression coefficients for Eqs. (8)–(11), which utilize data from the multiphysics simulation to predict the molten mass and operating temperature for
an MRE reactor.
Regolith m1 m2 m3 m4 m5 m6 t1 t2 t3 t4 t5 t6 t7
Highlands 11.0 5.31 0.160 4.99 0.343 0.435 1525 5387 0.351 0.0289 0.0105 0.215 5.09
Mare 9.73 3.98 0.226 5.60 0.500 0.485 1598 2200 0.593 0.0434 0.0390 0.419 1.54
S.S. Schreiner et al. / Advances in Space Research 57 (2016) 1585–1603 1595

5.5. Determining wall thickness and MLI design heat transfer equation for the predicted heat loss from the
reactor (Q_ heat-loss ), discussed Section 6:
Although the thermal conductivity of the reactor wall is !1=4
treated as a design variable in the multiphysics simulation Q_ heat-loss þ Ar1 ðT 1 Þ4
T out ¼ ð18Þ
and design methodology described above (doing so greatly Ar
decreased the time required to simulate a large design
space), the true thermal design variables are the wall mate- The MRE model leverages Eqs. (13)–(18) to calculate
rial, wall thickness and number of layers of multilayer insu- the required thermal resistance as a design target. This
lation (MLI). The MRE reactor model designs a reactor design target is then translated into an equivalent reactor
with real refractory and insulation materials such that its with MLI and real insulation and refractory materials
thermal resistance is equivalent to the thermal resistance according to the following methodology. First a mini-
of the reactor from the multiphysics simulation, which has mum wall thickness is set as a user input. Gmitter
a 3 cm thick wall, an outer wall emissivity of 0.05, the ther- (2008) utilized two 1 cm thick crucible walls, so a mini-
mal conductivity prescribed in Eq. (13), and the maximum mum wall thickness of 2 cm was used for this study.
diameter prescribed from Eq. (11). The thermal resistance Next, layers of MLI are added in an attempt to meet
of the reactor includes that of the reactor wall (Rcylinder ) the required thermal resistance design target. If this fails
and that of the radiation from the reactor wall through lay- to achieve the required thermal resistance (the effect of
ers of MLI to a distant ambient temperature (Rradiation ) adding layers of MLI has diminishing returns), the wall
(Kreith et al., 2011): thickness is also increased to achieve the thermal resis-
tance design target.
Rtotal ¼ Rcylinder þ Rradiation ; ð14Þ
The mass of the MLI was taken to be 0.02 kg/m2 per
where: layer (Plachta and Kittel, 2002), which includes a factor
1 of two to be conservative. The refractory material was cho-
Rcylinder ¼  2
 ; ð15Þ sen to be CW Hi Al Kastite C (high alumina)TM,5 with a
ð Din
þDxÞ ðHin þ2DxÞ
 density of 2.8 g/cm3 and a thermal conductivity of 1.4 W/
2pk eff 2
Dx
þ
ln 1þ2Dx
D in m-K. Aeroguard HDTM,6 was selected as the insulation
   1
Rradiation ¼  r T 2out þ T 21 ðT out þ T 1 Þ A ; ð16Þ material, for its excellent thermal properties at high tem-
peratures (a wall thermal conductivity of 0.02 to 0.06 W/
where k eff and Dx are the effective thermal conductivity and m-K depending upon temperature) and low density of
thickness, respectively, of the combined refractory and 0.2 g/cm3. Titanium was selected as the outer structural
insulation layers, Din is the reactor inner diameter, H in is material for its tolerance of high temperatures. Future
the reactor inner height (set to 1.5 times the electrode sep- work can examine additional materials to further optimize
aration), T out is the temperature between the reactor surface MRE reactor design.
and MLI, T 1 is the ambient temperature, A is the reactor A thickness ratio of 3:1 (insulation:refractory thick-
outer surface area, r is the Stefan–Boltzmann constant, ness) was used on the reactor bottom and side walls
and  is the effective emissivity of the layers of MLI and 6:1 was used on the reactor top, as the contact
(Donabedian et al., 2002): between molten regolith and the roof of the reactor will
 1 be minimal. The structural thickness is set such that the
 2N 1 1
 ¼ N 1þ þ ; ð17Þ hoop stress resulting from the operating pressure of
MLI surf 1
101.3 kPa is less than the yield stress of the material with
where N is the number of MLI layers, MLI is the emissivity of a safety factor of 3.5.
each MLI layer, taken to be 0.03 for Mylar (Donabedian
et al., 2002), and surf and 1 are the emissivities of the 6. Estimating reactor mass and power
reactor outer surface and environment, respectively (taken
to be 1.0). Although Donabedian et al. (2002) assert that The MRE reactor model estimates reactor mass using
the emissivity in Eq. (17) is a dramatic underestimate, recent the geometry of the reactor described above. With the
work by Hatakenaka et al. (2013) demonstrated a new pin diameter, height, and thickness of the structural, insula-
attachment method for MLI that allows for hardware tion, and refractory layers determined, the mass of the
performance to approach the theoretical predictions of three cylindrical reactor shells can be calculated.
Eq. (17). The mass of the anode and cathode are also estimated
Although the ambient temperature will dramatically using the simple shaft and plate geometry shown in
vary through the lunar cycle, it was taken to be 298 K after Fig. 1. Iridium and molybdenum were chosen for the anode
an initial analysis revealed low sensitivity to this parameter. and cathode materials, respectively, as they have significant
Varying the ambient temperature between 3 K and 400 K
resulted in less than a 2% change in the reactor thermal 5
http://www.alliedmineral.com/.
performance compared to the 298 K case, to first order. 6
http://www.microthermgroup.com/landingpage/assets/TDS_AERO-
T out in Eqs. (15) and (16) was calculated using the radiative GUARD_V1-EN.pdf.
1596 S.S. Schreiner et al. / Advances in Space Research 57 (2016) 1585–1603

experimental heritage operating with lunar simulant feed- Table 3


stock (Paramore, 2010). The diameter of the cathode plate The regression coefficients for Eq. (20), which predicts the expected heat
loss from an MRE reactor based off of data from the multiphysics
is set to 15% of the reactor diameter, after an initial anal- simulation.
ysis revealed that this ratio provided a good balance
Regolith h1 h2 h3 h4 h5 h6
between larger molten masses and higher operating tem-
peratures. The anode diameter is designed to be 19% less Highlands 1.59  106
8.29 1.08 1.15 222 0.0357
Mare 1.76  106 8.84 1.17 1.14 189 0.0345
than the cathode diameter, using the same ratio as Sibille
and Dominguez (2012). The cathode and anode shafts
are designed to be long enough to pass through the reactor
wall. The diameters of the anode and cathode shafts are set
to 2 cm plus 251 th of the plate diameter.
The total reactor power is estimated using the sum of
four different terms:
P total ¼ Q_ regheatup þ P ðDGÞ þ Q_ endo þ Q_ heat-loss ð19Þ
The power required to heat the regolith from ambient tem-
perature (298 K) to the operating temperature (Q_ regheatup ),
is calculated using the specific heat and latent heat of melt-
ing for lunar regolith from Schreiner et al. (2016). The heat
required to offset the endothermic electrolysis reaction
(Q_ endo ) is calculated from the difference between the
Enthalpy of Formation and Gibbs Free Energy provided
in Schreiner et al. (2016). The power required to perform
the electrolysis chemical reaction is calculated from the
Gibbs Free Energy data from Schreiner et al. (2016).
The radiative heat loss to the environment (Q_ heat-loss ) is
predicted using regression equations that were fit to the
data from the multiphysics simulation:
0  1
h2
exp Dþh
Q_ heat-loss ¼ h1 @ A þ h5 k wall ð20Þ
3

h4 þ kwall1h6

where k wall is given in Eq. (13), D is the diameter of the


reactor given by Eq. (11) and hi are regression coefficients
given in Table 3, which depend upon the regolith type.
Fig. 8 shows data generated by the multiphysics simula-
tion (dashed lines with data points) and the predictions of
Eq. (20) (solid lines). Although it may seem counterintu-
itive that reactor current does not significantly affect the
heat loss, this is because the electrode separation decreases
as current increases to maintain the joule heated cold wall
constraint as shown in Fig. 6. The MRE model shows that,
for a fixed diameter and wall conductivity, the thermal
topology required by the joule heated cold wall condition
imposes a certain heat loss on the reactor which does not
Fig. 8. The reactor heat loss does not significantly depend on current
significantly depend on the current.
(top), but is affected by wall thermal conductivity (bottom) when designing
Conversely, reactor heat loss does depend significantly on the cutoff lines. A nonlinear regression model (solid lines) was fit to the
on the wall thermal conductivity. Lower wall thermal data from a multiphysics simulation (dots with dashed lines) with
conductivities intuitively result in less heat loss, as a R2 = 0.997, RMSE = 0.51 kW.
direct result of the associated decrease in the required
electrode separation observed in Fig. 6. Fig. 8 demon- 7. Results: MRE performance and design trends
strates that by increasing the insulation of the reactor wall,
the electrode separation required to maintain joule heated 7.1. Oxygen extraction yield and current efficiency
cold wall operation decreases, resulting in a decrease in
resistive heating inside the reactor and therefore radiative The oxygen extraction yield (kO2 ) is the kilograms of
heat loss. oxygen that can be extracted per kilogram regolith, as
S.S. Schreiner et al. / Advances in Space Research 57 (2016) 1585–1603 1597

defined in Eq. (2). Fig. 9 shows how kO2 increases with near 2200 K is due to MgO also being electrolyzed, with
operating temperature for various types of lunar regolith. a low current efficiency of 50% (see Section 3).
As discussed in Section 2, higher operating temperatures The current efficiency estimates presented here are lower
enable an MRE reactor to electrolyze more of the oxides than many numbers in the literature, which are often
in lunar regolith before the melting temperature of the mol- around 90–100% (Sibille et al., 2010; Vai et al., 2010;
ten regolith core approaches the operating temperature, Sirk et al., 2010). Those estimates are for melts without
risking core solidification and reactor shutdown. Thus by FeO, TiO2, K2O, Na2O, which is why dramatically higher
raising the operating temperature, one can extract more current efficiencies are measured. Realistic current efficien-
oxygen per kilogram regolith and increase kO2 . Using com- cies over an entire batch of raw regolith, in which FeO,
position data from Schreiner et al. (2016), the maximum TiO2, and other oxides may be reduced, will be lower.
value for kO2 is 0.454 for Highlands, 0.417 for High-Ti
Mare, and 0.423 for Low-Ti Mare regolith. 7.2. The effect of operating temperature
For Highlands regolith, kO2 exhibits a sharp rise around
2000 K, at which point the reactor can reduce all of the Fig. 10 shows the how operating temperature affects
SiO2 (see Fig. 3: the SiO2 peak occurs around 1950 K, add- reactor mass (left) and power (right) over a range of oxygen
ing a 50 K safety margin yields 2000 K). For Mare regolith, production levels for High-Ti Mare regolith. Higher oper-
a jump in kO2 occurs around 2200 K. This jump is not due ating temperatures increase the oxygen extraction yield (see
to electrolyzing all of the SiO2, but is rather due to a rever- Fig. 9), which reduces the regolith throughput requirement
sal in the electrolysis order of SiO2 and MgO. At tempera- and reactor size. Higher operating temperatures also result
tures above 2200 K, the Gibbs Free Energy for MgO rises in a larger temperature drop that must be accommodated
above that of SiO2 (Schreiner et al., 2016), meaning that before reaching the reactor wall and a lower density for
MgO will be preferentially reduced over SiO2. In reality, molten regolith (Schreiner et al., 2016). Although the latter
it is likely that both MgO and SiO2 will be concurrently two effects drive larger reactor designs, they are outweighed
reduced at temperatures circa 2200 K, but for simplicity, by the increase in oxygen extraction yield and subsequent
the MRE reactor model assumes that oxide species are dis- decrease in reactor size shown in the left plot in Fig. 10.
cretely reduced in the order of increasingly negative Gibbs At lower temperatures (1850–2150 K), reactor power
Free Energy. increases with operating temperature due to the larger tem-
The right plot in Fig. 9 shows how the current efficiency perature drop, and therefore heat loss, between the reactor
(gI ) rises with temperature as more SiO2 (and other species) core and walls. At higher temperatures, this effect is out-
can be electrolyzed with a higher current efficiency than weighed by the steep increase in oxygen extraction yield
FeO, increasing the average current efficiency. For High- near 2200 K (see Fig. 9 for High-Ti Mare) and the increase
lands regolith, gI increases from 74% at an operating tem- in electrical conductivity which decreases the resistive heat-
perature of 1850 K to 82% above 2000 K. For Mare ing that must exit the reactor as heat loss.
regolith, gI increases from 57% at 1850 K to 72% at Fig. 10 also illustrates that, for a given batch time and
2200 K. The drop in current efficiency for Mare regolith design margin, a minimum achievable oxygen production

Fig. 9. The oxygen extraction yield (left) and current efficiency (right) of an MRE reactor generally rise with operating temperature, due to the fact that
more oxygen can be electrolytically extracted from lunar regolith at higher temperatures.
1598 S.S. Schreiner et al. / Advances in Space Research 57 (2016) 1585–1603

Fig. 10. Operating temperature affects both the mass (left) and power (right) for an MRE reactor. Each line represents a different operating temperature
and all designs have a design margin of 1.5 and a batch time of 8 h with High-Ti Mare feedstock.

level exists. Below a certain oxygen production threshold, electrode separation in order to maintain thermal equilib-
the reactor size becomes too small to maintain joule heated rium in the reactor. The left-hand plot in Fig. 12 demon-
cold wall operation. As operating temperature increases, it strates how increasing the design margin enables a range
becomes more difficult to design a small reactor and the of electrode separations that (1) generate enough heat to
minimum oxygen production threshold increases. For maintain the central molten core and (2) maintain the joule
High-Ti Mare with an operating temperature above heated cold wall operation (no molten regolith touches the
2200 K, this effect is outweighed by the dramatic increase reactor wall) by not generating excessive heat.
in oxygen extraction yield seen Fig. 9, and the minimum The multiphysics simulation described in Section 4 was
oxygen production threshold decreases with increasing leveraged to validate the range of electrode separations
operating temperature. enabled by a design margin greater than 1. The right side
of Fig. 12 shows reactor designs with the maximum (top)
and minimum (bottom) electrode separations for an oxy-
7.3. The effect of design margin gen production of 4000 kg O2/year with a design margin
of 1.2. The multiphysics simulation confirmed that both
The trends in operating temperature presented above reactor designs can support joule heated cold wall opera-
are for reactor designs with a fixed design margin of 1.5. tion by insulating the central molten core from the reactor
This means that the wall thermal conductivity is designed walls with a layer of solid regolith.
such that the maximum diameter set by the operating tem-
perature is 1.5 times the minimum diameter set by the mol-
ten mass requirement (see Section 5.4). Fig. 11 7.4. Sensitivity to regolith type/composition
demonstrates that increasing the design margin results in
higher reactor mass and power, and increases the maxi- The MRE reactor model has a strong dependence on
mum production rate for a single reactor. The maximum regolith type. Of the regolith material properties presented
production levels for design margins of 1.0 and 1.1 are in Schreiner et al. (2016), composition, density, specific
5000 kg O2/year and 9500 kg O2/year, respectively. Above heat, electrical conductivity, current efficiency, latent heat
these production rates, the MRE reactor model is unable to of melting, endothermic power, and electrolysis power all
produce a feasible reactor design which could sustain joule depend upon regolith composition. Furthermore, different
heated cold wall operation. sets of multiphysics data were generated for Highlands
Perhaps the most important effect of design margin is on and Mare regolith (the differences between High-Ti and
the range of acceptable electrode separation values. The Low-Ti Mare were not significant enough to warrant sepa-
design margin controls the acceptable bounds on reactor rate runs with the multiphysics simulation). This resulted in
diameter and because there is a one-to-one mapping separate sets of regression coefficients for the appropriate
between reactor diameter and electrode separation, this electrode separation (Table 1), molten mass and operating
also translates into bounds on the electrode separation temperature (Table 2), and heat loss (Table 3).
(see Section 5.3). As described in Section 5.4, during reac- Fig. 13 shows how the regolith type affects the mass and
tor operation a control system will need to regulate the power of an MRE reactor for operating temperatures of
S.S. Schreiner et al. / Advances in Space Research 57 (2016) 1585–1603 1599

Fig. 11. Larger design margins enable a range of acceptable electrode separations, but this comes at the cost of increased reactor mass (left) and power
(right). Larger design margins also increase the maximum production level for a reactor. All data is for an operating temperature of 1850 K and a batch
time of 18 h with High-Ti Mare feedstock.

Fig. 12. (left) Increasing the design margin opens up a larger range of feasible electrode separation values to enable operational flexibility. (right) At a
production level of 4000 kg O2/year, a design margin of 1.2 enables an acceptable electrode separation range of 1.9 cm to 3.8 cm, validated using the
multiphysics simulation.

1850 K (top row), 2000 K (middle row), and 2300 K (bot- requirement higher. Reactor mass remains relatively simi-
tom row). From these results, it appears that the general lar between Highlands and Mare because Highlands rego-
increase in regolith liquidus temperature during electrolysis lith has a lower electrical conductivity (Schreiner et al.,
(see Fig. 3) plays a large role in the regolith type dependence. 2016), which increases resistive heating in the central core
That is, the type of regolith strongly affects how much oxy- and requires less insulation around the reactor wall.
gen can be extracted from lunar regolith before having to
stop to avoid solidification of the central molten core. 7.5. Molten metal production
At 1850 K (top row of Fig. 13), one can extract less oxy-
gen per kilogram Highlands regolith compared to Mare The trend in liquidus temperature reverses around
regolith so Highlands reactors must process more regolith, 2000 K (middle row of Fig. 13), at which point an MRE
which increases reactor power. Furthermore, Highlands reactor can now extract more oxygen from Highlands rego-
regolith has more SiO2, which must be electrolyzed at a lith compared to Mare. This results in a dramatic decrease
higher voltage than FeO and subsequently drives the power in reactor mass (33%) and power (60%) for Highlands reac-
1600 S.S. Schreiner et al. / Advances in Space Research 57 (2016) 1585–1603

Fig. 13. The type of regolith feedstock strongly affects reactor mass and power at midrange operating temperatures circa 2000 K (middle row), where
Highlands reactors are significantly less massive and require less power than Mare reactors. At the low operating temperatures circa 1850 K (top row), and
high operating temperatures circa 2300 K (bottom row), the affect of regolith type is reduced, yet Highlands reactors are slightly less massive at high
production levels and require slightly more power across all production levels.
S.S. Schreiner et al. / Advances in Space Research 57 (2016) 1585–1603 1601

regolith, while reactors processing Highlands regolith can


produce more silicon and aluminum per kilogram regolith.
The data in Fig. 14 is for an oxygen production level of
10,000 kg O2/year. At higher operating temperatures
(>2000 K for Highlands and >2300 K for Mare), the
cumulative amount of metal produced is almost equal to
the oxygen production level. It is important to note that
the metals produced via MRE will alloy at the bottom of
the reactor. Post-reactor processing will be required if the
metals are to be separated, though this processing should
be feasible without significant additional energy expendi-
ture, as the metal alloy will already be molten.

8. Conclusions

We have developed a parametric sizing model for a Mol-


ten Regolith Electrolysis reactor to produce oxygen and
molten metals from lunar regolith. The reactor model is
grounded in a foundation of regolith material property
models presented in Schreiner et al. (2016), which are val-
idated using data from Apollo samples and lunar simu-
lants. The oxygen production level drives the regolith
processing requirement and reactor current. A multiphysics
simulation of an MRE reactor is developed and utilized to
generate a database of reactor designs and performance
characteristics. A novel design methodology is developed
that leverages this database to parametrically design
MRE reactors that can meet the required regolith process-
ing rate and reactor current. The predicted mass and power
of an MRE reactor are presented over a range of oxygen
production levels and the effects of operating temperature,
regolith type, and design margin (degree of operational
flexibility) are explored.
Our model predicts that an MRE reactor is able to
Fig. 14. The amount of metal produced by an MRE reactor operating on achieve oxygen extraction yields (kilograms of oxygen per
Mare (top) and Highlands (bottom) regolith producing 10 mT O2/yr. As kilogram regolith) on the order of 0.15 at low operating
operating temperature increases, more oxides can be reduced to produce temperatures and 0.375 at high operating temperatures.
more molten metal and leave less leftover slag.
We also observe that increasing the operating temperature
increases the predicted current efficiency, from 72.5% at
tors compared to Mare for operating temperatures circa low temperatures to 82% at higher temperatures for High-
2000 K. At operating temperatures above 2300 K, reactor lands regolith. For Mare regolith, predicted current effi-
mass and power become similar for both Highlands and ciencies start around 57.5% at low temperatures and rise
Mare reactors because the oxygen extraction yield is simi- to 70% at higher temperatures.
lar in that regime. Higher operating temperatures also appear to always
Fig. 14 illustrates the estimated metal produced from an reduce reactor mass and sometimes reduce power
MRE reactor, as discussed in Section 2, as a function of (Fig. 10), but this effect has a strong dependence on rego-
operating temperature for Mare (left) and Highlands lith type (Fig. 13). Prior to this work it was unclear whether
(right) regolith. As temperature increases, the metal pro- higher operating temperatures would indeed result in smal-
duction increases while the mass of leftover slag decreases ler reactor sizes due to the mixed effect of operating tem-
by approximately 35% from 1850 K to 2300 K (for Mare) perature on regolith material properties and reactor
or 2000 K (for Highlands). Reactors processing Highlands thermal topology. Higher operating temperatures also
regolith can produce aluminum at lower operating enable MRE reactors to produce more molten metal per
temperatures because the liquidus temperature peak occurs kilogram regolith and open up the possibility of producing
around 2000 K for Highlands regolith compared to 2250 K aluminum and titanium.
for Mare (see Fig. 3). Due to differences in regolith compo- Design margins greater than 1.0 enable a range of feasi-
sition (Schreiner et al., 2016), reactors processing Mare ble electrode separations (Fig. 12), which allows the reactor
regolith can produce more iron and titanium per kilogram thermal behavior to be controlled by varying electrode sep-
1602 S.S. Schreiner et al. / Advances in Space Research 57 (2016) 1585–1603

aration and resistive heating during operation. This opera- Balasubramaniam, R., Gokoglu, S., Hegde, U., 2010. The reduction of
tional flexibility comes at the cost of increased reactor mass lunar regolith by carbothermal processing using methane. Int. J.
Miner. Process. 96 (1), 54–61.
and power. Increasing the design margin also expands the Carr, B., 1963. Recovery of water or oxygen by reduction of lunar rock.
maximum production level of a single reactor (Fig. 11). AIAA J. 1 (4), 921–924.
Regolith type (and therefore composition) also has an Carroll, W.F., 1983. Research on the Use of Space Resources. JPL
impact on MRE reactor design and performance Publication, 83-36.
(Fig. 13). For midrange operating temperatures circa Chen, G.Z., Fray, D.J., Farthing, T.W., 2000. Direct electrochemical
reduction of titanium dioxide to titanium in molten calcium chloride.
2000 K, Highlands reactors are 33% less massive and Nature 407 (6802), 361–364.
require 60% less power than Mare reactors. At the low Chepko, A., 2009. Technology selection and architecture optimization of
and high ends of the operating temperature design space, in-situ resource utilization systems (PhD thesis). Massachusetts Insti-
the effect of regolith type becomes less significant. Never- tute of Technology.
theless, in these regimes Highlands reactors are marginally Colson, R.O., Haskin, L.A., 1990. Lunar oxygen and metal for use in
near-earth space: magma electrolysis. In: NASA Space Engineering
less massive at higher production levels and require slightly Research Center for Utilization of Local Planetary Resources.
more power across all production levels. Although MRE Curreri, P., Ethridge, E., Hudson, S., Miller, T., Grugel, R., Sen, S.,
reactors appear to generally perform slightly better with Sadoway, D., 2006. Process demonstration for lunar in situ resource
Highlands regolith, which covers 84% of the lunar nearside utilizationmolten oxide electrolysis. NASA Marshall Space Flight
and 99% of the lunar farside (Heiken et al., 1991), their per- Center. MSFC Independent Research and Development Project.
Project No. 5-81.
formance does not degrade significantly with Mare regolith. Diaz, J., Ruiz, B., Blair, B., Harsch, M., Duke, M., Parrish, C., Lueck, D.,
The MRE model provides initial evidence that MRE is rea- Mueller, R., Whitlow, J., 2005. Space transportation architectures and
sonably feedstock insensitive, allowing it to support surface refueling for lunar and interplanetary travel and exploration (starlite).
exploration across a large variety of lunar locations. Technical Report NASA Grant NAG9-1535, Center for the Commer-
Our results indicate that MRE reactors scale reasonably cial Applications of Combustion in Space, Colorado School of Mines.
Donabedian, M., Gilmore, D.G., Stultz, J., Tsuyuki, G.T., Lin, E.I., 2002.
with production level and that operational flexibility can Satellite Thermal Control Handbook: Fundamental Technologies.
indeed be parametrically designed into a reactor. We Duke, M.B., Diaz, J., Blair, B.R., Oderman, M., Vaucher, M., 2003.
emphasize that the mass and power results presented here Architecture studies for commercial production of propellants from
are for an unoptimized MRE reactor. Future work can the lunar poles. In: AIP Conference Proceedings. AIP Publishing, pp.
evaluate the optimization of MRE reactor design variables 1219–1226.
Faraday, M., 1834. Xxv. Experimental researches in electricity – seventh
in light of a more complete ISRU system, including a series. The London and Edinburgh Philosophical Magazine and
power system and the oxygen liquefaction and storage sys- Journal of Science, 5(27):161–181.
tem. This MRE reactor model will enable the quantitative Gerogiorgis, D., Ydstie, B., 2005. Multiphysics CFD modelling for design
comparison of MRE to other lunar oxygen production and simulation of a multiphase chemical reactor. Chem. Eng. Res.
methods, as well as the quantitative analysis of the feasibil- Des. 83 (6), 603–610.
Gibson, M.A., Knudsen, C.W., 1985. Lunar oxygen production from
ity of MRE in the broader context of space exploration ilmenite. In: Lunar bases and space activities of the 21st century, pp.
missions utilizing lunar ISRU. 543–550.
Gmitter, A.J., 2008. The influence of inert anode material and electrolyte
Acknowledgment composition on the electrochemical production of oxygen from molten
oxides (PhD thesis). Massachusetts Institute of Technology.
Haskin, L.A., Colson, R.O., Lindstrom, D.J., Lewis, R.H., Semkow, K.
This work was supported by a NASA Space Technology W., 1992. Electrolytic smelting of lunar rock for oxygen, iron, and
Research Fellowship (Grant #NNX13AL76H). Any opin- silicon. In: Lunar Bases and Space Activities of the 21st Century, vol.
ions, findings, and conclusions or recommendations 1, pp. 411–422.
expressed in this material are those of the author and do Hatakenaka, R., Miyakita, T., Sugita, H., Saitoh, M., Hirai, T., 2013.
not necessarily reflect the views of NASA. Thermal performance and practical utility of a mli blanket using
plastic pins for space use. In: 43rd International Conference on
Environmental Systems.
References Hegde, U., Balasubramaniam, R., Gokoglu, S., 2009. Development and
validation of a model for hydrogen reduction of jsc-1a. In: 47th
Aiken, R.H., 1906. Process of Making Iron from the Ore. US Patent Aerospace Sciences Meeting. AIAA, NASA/TM2009-215618.
816,142. Heiken, G.H., Vaniman, D.T., French, B.M., 1991. The Lunar source-
Allanore, A., Yin, L., Sadoway, D.R., 2013. A new anode material for oxygen book: A user’s guide to the Moon. CUP Archive.
evolution in molten oxide electrolysis. Nature 497 (7449), 353–356. Inman, D., White, S., 1978. The production of refractory metals by the
Allibert, M., Gaye, H., Geiseler, J., Janke, D., Keene, B.J., Kirner, D., electrolysis of molten salts; design factors and limitations. J. Appl.
Kowalski, M., Lehmann, J., Mills, K.C., Neuschutz, D., Parra, R., Electrochem. 8 (5), 375–390.
Saint-Jours, C., Spencer, P.J., Susa, M., Tmar, M., Woermann, E., Jarrett, N., Das, S.K., Haupin, W.E., 1980. Extraction of oxygen and
1995. Slag Atlas. Verlag Stahleisen, GmbH. metals from lunar ores. Space Sol. Power Rev. U.S. 1 (4).
Altenberg, B., 1990. Processing lunar in-situ resources. Technical report, Joosten, B.K., Guerra, L.A., 1993. Early lunar resource utilization: a key
Bechtel Group, Inc., Technical Research and Development Project Job to human exploration. In: AIAA Space Programs and Technologies
90634-002. Conference. AIAA, 93-4784.
Arndt, J., Flad, K., Feth, M., 1979. Radiative cooling experiments on Kar, P., Evans, J.W., 2008. A model for the electrochemical reduction of
lunar glass analogues. In: Lunar and Planetary Science Conference metal oxides in molten salt electrolytes. Electrochim. Acta 54 (2), 835–
Proceedings, vol. 10, pp. 355–373. 843.
S.S. Schreiner et al. / Advances in Space Research 57 (2016) 1585–1603 1603

Keller, R., Taberaux, A.T., 1991. Electrolysis of lunar resources in molten Senior, C.L., 1993. Lunar oxygen production by pyrolysis. In: Lewis, J.S.,
salt. Resources of Near-Earth Space: Proc. Second Annual Symp. UA/ Matthews, M.S., Guerrieri, M.L. (Eds.), Resources of Near-Earth
NASA SERC. Space. University of Arizona Press, Tucson, AZ, pp. 179–197.
Kennedy, M.W., 2012. Electric slag furnace dimensioning. In: Interna- Sherwood, B., Woodcock, G.R., 1993. Cost and benefits of lunar oxygen:
tional Smelting Technology Symposium: Incorporating the 6th economics, engineering, and operations. In: Lewis, J.S., Matthews, M.
Advances in Sulfide Smelting Symposium. John Wiley & Sons Inc, S., Guerrieri, M.L. (Eds.), Resources of Near-Earth Space. University
pp. 279–290. of Arizona Press, Tucson, AZ, pp. 199–227.
Kesterke, D.G., 1971. Electrowinning oxygen from silicate rocks. Tech- Sibille, L., Dominguez, J.A., 2012. Joule-heated molten regolith electrol-
nical report, U.S. Dept. of the Interior, Bureau of Mines. Report of ysis reactor concepts for oxygen and metals production on the moon
Investigations 7587. and mars. In: 50th AIAA Aerospace Sciences Meeting including the
Kim, H., Paramore, J., Allanore, A., Sadoway, D.R., 2011. Electrolysis of New Horizons Forum and Aerospace Exposition. AIAA, 2012-0639.
molten iron oxide with an iridium anode: the role of electrolyte Sibille, L., Sadoway, D.R., Sirk, A., Tripathy, P., Melendez, O., Standish,
basicity. J. Electrochem. Soc. 158 (10), E101–E105. E., Dominguez, J.A., Stefanescu, D.M., Curreri, P.A., Poizeau, S.,
Kreith, F., Manglik, R., Bohn, M., 2011. Principles of Heat Transfer,, 2009. Recent advances in scale-up development of molten regolith
Cengage Learning. electrolysis for oxygen production in support of a lunar base. In: 47th
Landis, G.A., 2005. Materials refining for solar array production on the AIAA Aerospace Sciences Meeting. American Institute of Aeronautics
moon. NASA Technical Memorandum. NASA/TM2005-214014. and Astronautics (AIAA).
Lindstrom, D.J., Haskin, L.A., 1979. Electrochemistry of lunar rocks. In: Sibille, L., Sadoway, D., Tripathy, P., Standish, E., Sirk, A., Melendez,
Fourth Princeton/AIAA Conference on Space Manufacturing Facil- O., Stefanescu, D., 2010. Performance testing of molten regolith
ities, pp. 79–1380. electrolysis with transfer of molten material for the production of
Lucey, P.G., Blewett, D.T., Jolliff, B.L., 2000. Lunar iron and titanium oxygen and metals on the moon. In: AIAA: 3rd Symposium on Space
abundance algorithms based on final processing of clementine ultra- Resource Utilization.
violet–visible images. J. Geophys. Res. Planet. (1991–2012) 105 (E8), Simon, M.C., 1985. A parametric analysis of lunar oxygen production. In:
20297–20305. Lunar Bases and Space Activities of the 21st Century.
Mason, L., 1992. Beneficiation and comminution circuit for the produc- Sirk, A.H., Sadoway, D.R., Sibille, L., 2010. Direct electrolysis of molten
tion of lunar liquid oxygen. In: Engineering, Construction, and lunar regolith for the production of oxygen and metals on the moon.
Operations in Space III, pp. 1139–1149. ECS Trans. 28 (6), 367–373.
Paramore, J.D., 2010. Candidate anode materials for iron production by Standish, E., 2010. Design of a molten materials handling device for
molten oxide electrolysis (PhD thesis). Massachusetts Institute of support of molten regolith electrolysis (PhD thesis). The Ohio State
Technology. University.
Plachta, D., Kittel, P., 2002. An updated zero boil-off cryogenic propellant Stoeser, D., Rickman, D., Wilson, S., 2010. Design and Specifications for
storage analysis applied to upper stages or depots in an leo environ- the Highland Regolith Prototype Simulants NU-LHT-1M and-2M.
ment. In: 38th AIAA/ASME/SAE/ASEE Joint Propulsion Conference Citeseer.
and Exhibit. AIAA, 2002-3589. Taylor, L.A., Carrier III, W.D., 1993. Oxygen production on the moon:
Poizeau, S., Sadoway, D.R., 2011. Towards a design tool for self-heated an overview and evaluation. In: Lewis, J.S., Matthews, M.S., Guer-
cells producing liquid metal by electrolysis. Light Metals 2011, 387–392. rieri, M.L. (Eds.), Resources of Near-Earth Space. University of
Sacksteder, K.R., Sanders, G.B., 2007. In-situ resource utilization for Arizona Press, Tucson, AZ, pp. 69–108.
lunar and mars exploration. In: The 5th AIAA Aerospace Sciences Teeple, B.S., 1994. Feasibility of producing lunar liquid oxygen (PhD
Meeting and Exhibit. AIAA, 2007-345. thesis). Massachusetts Institute of Technology.
Schreiner, S.S., 2015. Molten regolith electrolysis reactor modeling and Turner, M.J., 2008. Rocket and Spacecraft Propulsion: Principles, Practice
optimization of in-situ resource utilization systems (Master’s thesis). and New Developments. Springer Science & Business Media.
Massachusetts Institute of Technology. Vai, A., Yurko, J., Wang, D., Sadoway, D., 2010. Molten oxide
Schreiner, S.S., Sibille, L., Dominguez, J.A., Hoffman, J.A., Sanders, G. electrolysis for lunar oxygen generation using in situ resources. In:
B., Sirk, A.H., 2015. Development of a molten regolith electrolysis The Minerals, Metals, and Materials Society.
reactor model for lunar in-situ resource utilization. In: AIAA SciTech Wang, D., Gmitter, A.J., Sadoway, D.R., 2011. Production of oxygen gas
Conference – 8th Symposium on Space Resource Utilization. and liquid metal by electrochemical decomposition of molten iron
Schreiner, S.S., Dominguez, J.A., Sibille, L., Hoffman, J.A., 2016. oxide. J. Electrochem. Soc. 158 (6), 51–54.
Thermophysical property models for lunar regolith. Adv. Space Res. Yen, C.-A.F., 1977. Electrical conductivity in the FeO Fe2O3–Al2O3–
Schwandt, C., Hamilton, J.A., Fray, D.J., Crawford, I.A., 2012. The SiO2 system (PhD thesis). Massachusetts Institute of Technology.
production of oxygen and metal from lunar regolith. Planet. Space Sci.
74 (1), 49–56.

Você também pode gostar