You are on page 1of 25

Solar Energy xxx (2017) xxxxxx

Contents lists available at ScienceDirect

Solar Energy
journal homepage: www.elsevier.com/locate/solener

Solar fuels via chemical-looping reforming


Peter T. Krenzke a, Jesse R. Fosheim b, Jane H. Davidson b,
a
Physics & Engineering Department, Taylor University, 236 W. Reade Ave, Upland, IN 46989, United States
b
Department of Mechanical Engineering, University of Minnesota, 111 Church St. S.E., Minneapolis, MN 55455, United States

a r t i c l e i n f o a b s t r a c t

Article history: The potential of solar thermal chemical-looping reforming for efficient and sustainable co-production of
Received 26 February 2017 synthesis gas and hydrogen is discussed. In an endothermic partial oxidation step, methane reacts with
Received in revised form 20 May 2017 oxygen released from a metal oxide to produce hydrogen and carbon monoxide. In an exothermic second
Accepted 29 May 2017
step, steam reacts with the reduced metal oxide to produce hydrogen. This review summarizes the pro-
Available online xxxx
cess and chemical thermodynamic foundations of solar chemical-looping reforming and provides a syn-
opsis of materials studies that reflect the state of knowledge in 2017. The challenges and opportunities
Keywords:
for future research and development are discussed.
Fuel
Solar thermochemical
2017 Published by Elsevier Ltd.
Metal oxide
Reforming

1. Introduction et al., 2001). Furthermore, safety measures must be implemented


to mitigate the risk of explosion (Said et al., 2016). An auxiliary
The development of processes to produce sustainable alterna- stream of ca. 3:1 H2:CO syngas is obtained via methane steam
tives to petroleum-derived fuels is imperative given the expanding reforming (R2) and converted by water-gas shift (R3) and pressure
global demand for liquid fuels. Hydrocarbon fuels are projected to swing adsorption to high purity hydrogen.
be important for transportation in the foreseeable future even with
CH4 H2 O ! 3H2 CO; DH 206 kJ=molCH4 R2
an expanding market for electric vehicles. In 2015, transportation
comprised 28% of total energy consumption in the U.S. with 92%
supplied by petroleum (EIA, 2016a). Globally, energy consumption CO H2 O ! CO2 H2 ; DH 41 kJ=molCH4 R3
in the transportation sector is predicted to increase 1.4% annually
The hydrogen is used to tailor the H2:CO ratio of the syngas stream
to 163.5 EJ (155 quadrillion Btu) in 2040 (EIA, 2016b).
to match the requirements for GTL processing.
One approach to meet the growing demand for liquid fuels pro-
Chemical-looping reforming (CLR) is a two-step thermochemi-
duced sustainably is to use solar energy to offset a portion of the
cal cycle to produce syngas and hydrogen (or carbon monoxide).
energy required for Gas-to-Liquid (GTL) processes. GTL processes
In the endothermic partial oxidation step (R4), methane reacts
convert a mixture of hydrogen and carbon monoxide (termed syn-
with oxygen released from a metal oxide (MOx) to produce syngas.
thesis gas or syngas) to liquid fuels, including high cetane diesel,
In the exothermic oxidizer splitting step (R5), H2O and/or CO2 react
high octane gasoline, and kerosene (jet fuel) (Wright et al., 2003;
with the reduced metal oxide to produce H2 and/or CO.
Wood et al., 2012). In conventional GTL, methane is partially oxi-
dized via (R1) to ca. 2:1 H2:CO syngas (Uhde, 1997; de Klerk, CH4 1=DdMOXdOX ! 1=DdMOXdRD 2H2 CO; DH > 0 R4
2011; Wilhelm et al., 2001; Said et al., 2016).
aH2 O 1  aCO2 1=DdMOXdRD
CH4 1=2O2 ! 2H2 CO; DH 36 kJ=molCH4 R1
! 1=DdMOXdOX aH2 1  aCO; DH < 0 R5
Reaction (R1) is implemented almost exclusively with high pur-
ity oxygen (Wilhelm et al., 2001; Said et al., 2016; Rostrup-Nielsen, In comparison to reduction of a metal oxide in an inert sweep
2002). Air separation is energy intensive and the oxygen plant can gas or sub-atmospheric pressure, the introduction of methane in
comprise up to 40% of the total cost of a syngas plant (Wilhelm reaction (R4) lowers the thermodynamic barrier to split water (or
carbon dioxide), such that the thermodynamics of CLR are those
Corresponding author. of methane reforming rather than the less favorable thermody-
E-mail addresses: peter_krenzke@taylor.edu (P.T. Krenzke), foshe002@umn.edu namics of thermolysis. The net products of CLR match those of
(J.R. Fosheim), jhd@umn.edu (J.H. Davidson). steam and dry reforming for steam fractions of a = 1 and a = 0,

http://dx.doi.org/10.1016/j.solener.2017.05.095
0038-092X/ 2017 Published by Elsevier Ltd.

Please cite this article in press as: Krenzke, P.T., et al. Solar fuels via chemical-looping reforming. Sol. Energy (2017), http://dx.doi.org/10.1016/j.
solener.2017.05.095
2 P.T. Krenzke et al. / Solar Energy xxx (2017) xxxxxx

Nomenclature

C solar concentration ratio g efficiency


F loss coefficient r Stefan-Boltzmann constant (W m2 K1)
h molar enthalpy (J mol1)  active site
HHV higher heating value (J mol1)
I direct normal irradiance (1000 W m2) Superscripts
ni number of moles of species (mol) chemisorbed species
n_ 0i mass specific molar species flow rate (mol s1 g1)
P pressure (bar) Subscripts
pO2 oxygen partial pressure (bar) 1 ambient conditions (1 bar, 298 K)
P1 total pressure (1 bar) 1, 2, . . . state
Q heat (J)
FC fuel cell
Q_ heat transfer rate (W) GAS energy associated with process gases
R ideal gas constant (8.314 J mol1 K1), CH4:O2 ratio i species
(Figs. 37) IN inlet of the reactive bed, entering the system
SC carbon selectivity
LOSS conduction/convection to ambient
SCO carbon monoxide selectivity OUT outlet of the reactive bed, exiting the system
SH2 hydrogen selectivity OX oxidation, oxidizer
t time (s)
RAD radiative emission
T temperature (K) REACTOR solar reactor, i.e. includes conduction, convection, and
V_ flow rate of gas per gram of oxygen carrier emission losses
(mL min1 g1) REJECT excess heat that is rejected to maintain isothermal oper-
W work (J)
ation
XCH4 methane conversion RD reduction
XCO2 carbon dioxide conversion SEP separation
XH2 O water conversion SOLAR solar
y mole fraction
S?E solar-to-electric conversion

Symbols Abbreviations
a steam fraction used in (R5) CLR chemical looping reforming
d nonstoichiometry
SCLR solar chemical looping reforming
DP pressure drop (bar) SSA specific surface area (m2 g1)
Dd change in nonstoichiometry

P
respectively. Decoupling the partial oxidation and oxidizer split- i ni;OUT HHVi
U 1
ting steps provides advantages over conventional reforming. Syn- nCH4;IN HHVCH4
gas produced via (R4) has a H2:CO ratio of 2:1 compared to 3:1
With complete conversion of methane and oxidizer to syngas,
and 1:1 for steam and dry reforming, respectively. The syngas from
the energetic upgrade factor is 128%. Combustion of the solar
(R5) can be tuned via the steam fraction, a, to obtain any H2:CO
upgraded syngas emits 59% of the carbon per kilojoule that would
ratio. Operating with a = 1 (i.e. water splitting) yields separate
be emitted in the combustion of syngas from conventional CLR.
streams of syngas and high purity H2. For CO2 splitting, the ener-
Solar-to-chemical efficiency is commonly used to project com-
getic and economic costs of providing pure CO2 will be high (Kim
mercial viability of a nascent solar thermochemical cycle such as
et al., 2011, 2012; Herron et al., 2015). This cost is one reason for
SCLR. Efficiency has been used traditionally as a surrogate for cost,
a preference for water splitting. To prevent catalyst deactivation
although the authors and the research community recognize that
by carbon deposition, conventional reformers must operate with
more complex factors including the value of competing technolo-
excess oxidizer (e.g.H2O:CH4  3), which increases energy require-
gies are important. The solar-to-chemical efficiency, given by Eq.
ments and lowers process efficiency (Simakov et al., 2015). CLR
(2), is defined as the net gain in the higher heating value (HHV)
does not require excess oxidizer because carbon deposition can
of the gases divided by the direct solar thermal power input to
be prevented by limiting the extent of reduction of the metal oxide
the reactor, Q SOLAR , and the solar thermal power required to pro-
(Otsuka et al., 1993, 1997, 1998; Cho et al., 2005). Conventionally,
vide parasitic work, W=gS!E , including that required for separa-
process heat for CLR is supplied by combusting up to 41% of the
tions, pumping, etc.
methane feedstock (Simakov et al., 2015), resulting in a 24% reduc-
tion in energy content compared to the feedstock. P
i ni;OUT  ni;IN HHVi
Feedstock utilization and CO2 emission can be improved by g 2
Q SOLAR W=gS!E
supplying process heat with concentrated solar energy, as illus-
trated in Fig. 1. Solar CLR (SCLR) is net endothermic. The chemical Based on reasonable assumptions for the parasitic energy
energy requirements for producing synthesis gas are identical to requirements and operation of a concentrated solar thermochemi-
the energy required for solar steam methane reforming (SMR) cal reactor, Krenzke et al. project an efficiency of SCLR of 54%. Key
or solar dry methane reforming (DMR). Solar energy is stored in assumptions in this projection are complete conversion of reac-
the products and the syngas has a higher energy content than tants to syngas and hydrogen in (R4) and (R5) and operation at
the methane feedstock. The energetic upgrade factor, U, given in 1273 K and 1000 suns solar input (1 sun equals 1000 W m2 con-
Eq. (1), is the ratio of the total higher heating value (HHV) of centration). Consistent with the majority of studies included in this
the products (species i) to the heating value of the methane review, the projection assumes isothermal operation of the cycle at
feedstock. atmospheric pressure.

Please cite this article in press as: Krenzke, P.T., et al. Solar fuels via chemical-looping reforming. Sol. Energy (2017), http://dx.doi.org/10.1016/j.
solener.2017.05.095
P.T. Krenzke et al. / Solar Energy xxx (2017) xxxxxx 3

1
nCH4 ;IN  8
CH4 H 2O XCH4 4  2SH2  SCO
To date, research on SCLR has focused primarily on process
thermodynamics (Krenzke et al., 2016; Steinfeld et al., 1996;
Forster, 2004; Krenzke and Davidson, 2014; Warren et al., 2017).
Solar Chemical Numerous material studies for CLR and limited studies for SCLR
Looping Reforming are relevant. Here we review the prior thermodynamic analyses
of SCLR, interpret the relevant experimental studies of metal oxi-
des as they relate to SCLR, and in conclusion recommend avenues
of future inquiry.

Separation Separation 2. Thermodynamics

In Section 2, we present the chemical thermodynamics under-


pinning SCLR, a review of prior thermodynamic models of the SCLR
Syngas H2 process, an expansion of the prior process thermodynamic analysis
H2:CO=2:1 by Krenzke et al. (2016), and projections of solar-to-chemical effi-
ciency. The expanded thermodynamic analysis accounts for the
work to extract syngas from the products of (R4). The efficiency
projections show the impact of key energy requirements on
GTL Fuel Cell efficiency.

2.1. Chemical thermodynamics


Liquid Fuels Work
2.1.1. Reduction
Fig. 1. Solar chemical looping reforming (SCLR). The process utilizes process heat The equilibrium product distribution for the partial oxidation of
from concentrated solar energy to upgrade CH4 and H2O via a metal oxide redox
methane (R1) at 1 bar is shown in Fig. 2. At 373 K, the products are
cycle. Ideally SCLR produces syngas with H2:CO = 2:1 and pure H2. For non-ideal
operation, separation processes require additional work input to extract the desired a mixture of water, carbon dioxide, solid carbon, and methane.
products. From 373 K to 1373 K, the oxygen mole fraction remains below
1016, which is substantially lower than could be achieved by vac-
uum pumping or inert gas sweeping. Syngas comprises more than
The primary challenge of obtaining commercially relevant 95% of the products above 1100 K and over 99% above 1223 K.
solar-to-chemical efficiencies with SCLR is achieving high methane The partial oxidation of methane (R1) and the equilibrium pro-
and oxidizer conversion and syngas selectivity (Krenzke et al., duct distribution in Fig. 2 require a 2:1 ratio of methane to oxygen
2016). In addition, deposition of solid carbon should be avoided (CH4:O2). In implementing CLR, the CH4:O2 ratio depends on the
to avoid deactivation of the metal oxide. The key metrics for the rate of metal oxide reduction because the metal oxide is the oxy-
conversion of methane to synthesis gas via reaction (R4) are gen source. Rate expressions for reduction of nonstoichiometric
methane conversion (XCH4 ), hydrogen selectivity (SH2 ), carbon oxides such as ceria depend on nonstoichiometry. Thus, as
monoxide selectivity (SCO ), and carbon selectivity (SC ). For reaction observed for ceria, the CH4:O2 ratio is not constant during reaction
(R5), a key metric is the oxidizer conversion (XOX ). (R4) (Krenzke et al., 2016).
XCH4  nCH4 ;IN  nCH4 ;OUT =nCH4 ;IN 3 The impact of deviation from CH4:O2 = 2:1 on the equilibrium
product distribution is shown in Fig. 3. (The CH4:O2 ratio in the fig-
SH2  nH2 ;OUT =2XCH4 nCH4 ;IN 4 ures is labeled R to improve the readability.) For low CH4:O2
ratios, e.g. 1:1, excess oxygen promotes complete combustion of
SCO  nCO;OUT =XCH4 nCH4 ;IN 5 methane. The presence of excess methane, e.g. CH4:O2 = 4:1,
reduces the amount of water and carbon dioxide. From 373 K to
SC  nC =XCH4 nCH4 ;IN 6 1373 K, unreacted methane and solid carbon are present. Carbon
deposition has been observed in experimental evaluations of CLR
XOX  nOX;IN  nOX;OUT =nOX;IN 7 in fixed bed reactors, most commonly when much of the available
oxygen from the metal oxide has been consumed in (R4) (Otsuka
The reactions (R4) and (R5) are expressed in terms of conversions
et al., 1993, 1997, 1998; Cho et al., 2005). Deposition of carbon
and selectivities in (R6) and (R7). Reaction (R6) is written assuming
on the oxide can block reaction sites. Under some conditions, the
that converted methane forms only hydrogen, water, carbon
issue can be mitigated by gasification of the carbon during the oxi-
monoxide, and carbon dioxide. Reaction (R7) is for water splitting.
dation step (R5) (Snoeck and Froment, 2003; Sang et al., 2014), but
1 1 in the case of water splitting, the hydrogen product stream is then
MOxdOX nCH4 ;IN CH4 ! MOxdRD 1  XCH4 nCH4 ;IN CH4
Dd Dd contaminated with carbon oxides.
h
XCH4 nCH4 ;IN 2SH2 H2 21  SH2 H2 O The methane conversion predicted by chemical equilibrium is
i shown in Fig. 4 for CH4:O2 from 1:2 to 8:1. For CH4:O2 = 1:2,
SCO CO 1  SCO CO2 R6 enough oxygen is present to completely combust the methane
leading to XCH4 = 1 from 373 K to 1373 K. For CH4:O2 > 2:1, the
1 1 1 1 - XH2 O reaction is methane-rich and the lack of oxygen results in lower
MOxdRD H2 O ! MOxdOX H2 H2 O R7 methane conversion.
Dd XH2 O Dd XH2 O
Hydrogen and carbon monoxide selectivities are shown in
The quantity of methane supplied is defined in Eq. (8) in terms of Fig. 5. The selectivity toward hydrogen improves as the CH4:O2
the conversion and selectivity definitions in Eqs. (3)(5). ratio is increased. The selectivity toward CO, however, is maxi-

Please cite this article in press as: Krenzke, P.T., et al. Solar fuels via chemical-looping reforming. Sol. Energy (2017), http://dx.doi.org/10.1016/j.
solener.2017.05.095
4 P.T. Krenzke et al. / Solar Energy xxx (2017) xxxxxx

1 1

0.9 R=0.5

0.8 0.8
H2

Methane Conversion
R=1
0.7
Mole Fraction

0.6
0.6
H 2O
0.5 R=2
0.4 CO
CH4
0.4

0.3 R=4
0.2 C
O2
CO2 0.2 R=8
0 0.1
373 573 773 973 1173 1373 373 573 773 973 1173 1373
Temperature [K] Temperature [K]

Fig. 2. Equilibrium product distribution of the partial oxidation of methane (CH4: Fig. 4. Methane conversion predicted by chemical equilibrium for CH4:O2 = 1:2 to
O2 = 2). 8:1. Partial oxidation of methane corresponds to CH4:O2 = 2:1. Methane combustion
corresponds to CH4:O2 = 2:1 for which XCH4 = 1. For CH4:O2 > 2:1, the reaction is
methane-rich.

mized at CH4:O2 = 2:1. For CH4:O2 < 2:1, carbon monoxide selectiv-
ity is depressed due to the formation of carbon dioxide. CH4: reaches equilibrium with the oxygen partial pressure plotted ver-
O2 > 2:1 promotes the formation of solid carbon. In comparison sus temperature in Fig. 6. The oxide is then exposed to steam,
to hydrogen selectivity, selectivity toward carbon monoxide is which provides the increase in oxygen partial pressure necessary
more sensitive to the CH4:O2 ratio. In practice, carbon formation for oxidation of the reduced metal oxide, leaving a mixture of
may be limited by kinetics or heat transfer (Snoeck et al., 1997a, hydrogen and steam. Possible compositions of the gas mixture
1997b) such that one may be able to use excess methane to achieve range from pure hydrogen to pure steam. Pure steam corresponds
high selectivities toward hydrogen and carbon monoxide. to XH2 O = 0 and the highest possible equilibrium oxygen partial
The link between oxide-independent and oxide-specific equilib-
pressure for the splitting reaction. In the limit as XH2 O approaches
rium predictions is the oxygen partial pressure, which is shown in
unity, corresponding to pure hydrogen, the oxygen partial pressure
Fig. 6 for CH4:O2 from 1:2 to 8:1. At 1 bar total pressure, the equi-
approaches zero.
librium oxygen partial pressure is the O2 mole fraction. The oxygen
The equilibrium oxygen partial pressure of (R7) is shown
partial pressure increases with increasing temperature. The oxygen
(dashed curves) as a function of temperature for oxidizer conver-
partial pressure is highest for CH4:O2 = 1:2 because there is suffi-
sions ranging from 0.01 to 0.99 in Fig. 7. The equilibrium oxygen
cient oxygen for complete combustion of the methane.
partial pressures, shown in Fig. 6 for (R6), are overlaid (solid
curves) to illustrate accessible isothermal cycles for which the
2.1.2. Oxidation equilibrium oxygen partial pressure is higher for (R7) than for
The ability of the reduced oxide to split water or carbon dioxide (R6). For example, the equilibrium oxygen partial pressure for
depends on its state following reaction with methane. Consider (R6) with CH4:O2 = 1:1 falls between the oxidizer conversion
isothermal operation of SCLR for water splitting. Oxygen is curves for XH2 O = 0.5 and 0.9. Therefore XH2 O = 0.5 is accessible with
removed from the oxide during reaction (R6) until the oxide CH4:O2 = 1:1, but XH2 O = 0.9 is not. For CH4:O2 = 1:2 in (R6),

1 1
R=1 R=4

0.8 0.8
H2
Mole Fraction
Mole Fraction

0.6 H 2O 0.6
CH4
H2

0.4 0.4 H 2O
CO2
CH4 C
0.2 0.2
CO
CO
CO2
C O2 O2
0 0
373 573 773 973 1173 1373 373 573 773 973 1173 1373
Temperature [K] Temperature [K]

Fig. 3. Equilibrium product distribution for the reaction of methane and oxygen for CH4:O2 = 1:1 and 4:1 (CH4:O2 is indicated by R). (Left) Excess oxygen is available with R :
= 1. (Right) Excess methane is supplied to the reaction (R = 4).

Please cite this article in press as: Krenzke, P.T., et al. Solar fuels via chemical-looping reforming. Sol. Energy (2017), http://dx.doi.org/10.1016/j.
solener.2017.05.095
P.T. Krenzke et al. / Solar Energy xxx (2017) xxxxxx 5

1 1
R=2
R=2
0.8 R=8 0.8 R=1

CO Selectivity
H2 Selectivity
0.6 R=4 0.6
R=1 R=4

0.4 0.4
R=8

0.2 0.2

R=0.5 R=0.5
0 0
373 573 773 973 1173 1373 373 573 773 973 1173 1373
Temperature [K] Temperature [K]

Fig. 5. Equilibrium selectivities toward hydrogen (left) and carbon monoxide (right) for CH4:O2 from 1:2 to 8:1.

100 100
R=0.5
R=0.5 -5
-5 10
10

10-10 10-10
R=1
R=1 2
-15
.01
2

pO

10-15 10
pO

O
=0
XH2 9
R=2 =0.
.5 X H 2O
10 -20 10 -20 =0
XH2 O
R=2

R=4
10-25 R=8 10-25 .99
=0
O
X H2

10 -30 10-30
873 973 1073 1173 1273 1373 873 973 1073 1173 1273 1373
Temperature [K] Temperature [K]

Fig. 6. Equilibrium oxygen partial pressure for CH4:O2 from 1:2 to 8:1. Fig. 7. Oxidation equilibrium mole fraction (dashed curves) as a function of
temperature for oxidizer conversions between 0.01 and 0.99. Equilibrium oxygen
mole fractions for (R6) are shown (solid curves) for comparison. For nearly
completely converted oxidizer, the equilibrium oxygen mole fraction is several
oxidizer conversion is less than 0.01. Thus, it is not possible to orders of magnitude lower than for low oxidizer conversion.
completely convert methane to water and carbon dioxide in (R6)
and to achieve high oxidizer conversion in (R7) with an isothermal
cycle. Very high oxidizer conversion, greater than 0.99, is possible
2012; Bader et al., 2013; Krenzke and Davidson, 2015; Furler
at 1273 K with CH4:O2 = 2:1 in (R6). This scenario matches the pre-
et al., 2012a; Ermanoski et al., 2013, 2014; Bulfin et al., 2015;
diction of Krenzke and Davidson that high quality syngas can be
Brendelberger et al., 2015; Ehrhart et al., 2016a, 2016b, 2016c;
produced with a 2:1 H2/CO ratio in (R6) along with achieving near
Venstrom et al., 2014; Muhich et al., 2013; Ermanoski, 2014; Lin
complete conversion of oxidizer to fuel in (R7) (Krenzke and
and Haussener, 2015), there are relatively few process thermody-
Davidson, 2014).
namic analyses of SCLR (Krenzke et al., 2016; Steinfeld et al.,
Operating (R7) with a lower temperature than (R6) improves
1996; Forster, 2004; Krenzke and Davidson, 2014; Warren et al.,
the thermodynamic favorability of water splitting. A potential ben-
2017). Even so, the analyses of SCLR exhibit substantial variation
efit of two-temperature cycling for solar fuel production via SCLR is
in approach, level of detail considered for the system, and defini-
improved oxidizer conversion, especially if (R6) is operated with
tion of efficiency. Here we summarize the findings of these studies
excess oxygen, i.e. CH4:O2 < 2:1. Unlike isothermal operation,
and extend the analysis of Krenzke et al. (2016) to include the
which can be evaluated independently of the metal oxide oxygen
energy requirements for separation of the partial oxidation prod-
carrier, the thermodynamics of two-temperature cycling are mate-
ucts. The results demonstrate the importance of methane conver-
rial specific.
sion, syngas selectivity, and oxidizer conversion on efficiency.
The prior thermodynamic studies consider both open (Krenzke
2.2. Process thermodynamics et al., 2016; Steinfeld et al., 1996; Krenzke and Davidson, 2014) and
closed systems (Steinfeld et al., 1996; Forster, 2004). The open sys-
Compared to the number of published thermodynamics analy- tem is shown in Fig. 8. Reactants (CH4 and H2O) and products (CH4,
ses of solar thermochemical metal oxide water and carbon dioxide H2, CO, H2O, and CO2) cross the boundary. The metal oxide is cycled
splitting (with inert sweep gas or vacuum reduction) (Lapp et al., between two reactors for (R6) and (R7). For closed systems, only

Please cite this article in press as: Krenzke, P.T., et al. Solar fuels via chemical-looping reforming. Sol. Energy (2017), http://dx.doi.org/10.1016/j.
solener.2017.05.095
6 P.T. Krenzke et al. / Solar Energy xxx (2017) xxxxxx

In an earlier study, Krenzke and Davidson (2014) used chemical


thermodynamics to predict equilibrium gas compositions for (R6)
and (R7), which are linked to conversions and selectivities as
shown in Section 2.1, and to project efficiency for a ceria-based
SCLR cycle. Later, Warren et al. published a similar process thermo-
dynamic analysis employing chemical equilibrium predictions
(Warren et al., 2017). Both studies enforce equilibrium between
fixed quantities of ceria and methane, but there are important dif-
ferences in how operating conditions are considered. Krenzke and
Davidson (2014) consider operating conditions which provide the
desired methane to oxygen ratio of CH4:O2 = 2:1. This ratio enables
98% methane conversion, hydrogen and carbon monoxide selectiv-
ities of 99%, and high efficiency for operating temperatures from
1223 to 1273 K. The 2:1 methane to oxygen ratio is achieved by
equating the ratio of ceria to methane to the change in nonstoi-
chiometry, Dd = dRD  dOX.
Fig. 8. Open system for solar chemical looping.
nCH4 =nceria Dd 11
heat and work, i.e. no chemical species, cross the system boundary. The nonstoichiometry of the reduced ceria (dRD) in equilibrium
To close the system, a fuel cell is added in which the products are with the products of the partial oxidation of methane varies with
recombined to form the reactants and extract electricity. reaction temperature. The oxidation nonstoichiometry (dOX) is
Steinfeld et al. studied the process thermodynamics of co- treated as a parameter to show its impact on oxidizer conversion
production of zinc and synthesis gas from zinc oxide and methane and efficiency. Accordingly, the ratio of methane to ceria varies
at 1250 K (Steinfeld et al., 1996). Efficiencies are reported using with temperature and oxidation nonstoichiometry in the range
Eq. (9). 0.1 < Dd < 0:245. In contrast, Warren et al. consider the reaction
between constant ratios of ceria and methane,
WFC
g 9
Q SOLAR nCH4 ;IN HHVCH4 nCH4 =nceria N 12

This definition differs from the solar-to-chemical efficiency given by for N = 0.25 and N = 1. With their approach, CH4:O2, varies with
Eq. (2) in that it uses a heat (QSOLAR + nCH4 ;IN HHVCH4 ) to electrical reaction temperature because of the accompanying variation in
work (WFC ) conversion. Energy requirements accounted for by the equilibrium nonstoichiometry. For N = 0.25, the methane to oxygen
efficiency calculations include irreversibilities in quenching, heat ratio varies from CH4:O2 = 70:1 at 923 K to CH4:O2 = 2:1 at 1323 K
transfer to and from the reactor, and chemical reactions. Pumping based on reported nonstoichiometries. For N = 1, CH4:O2 = 150:1
work was not considered. With heat recovery from the products, at 923 K and CH4:O2 = 8:1 at 1270 K. For CH4:O2 ratios greater than
the predicted open cycle efficiency is 43% for a hydrogenair fuel two, separation of syngas from the product stream to produce syn-
cell and 50% for a zinc-air fuel cell. gas for GTL would be necessary, but separation work is not consid-
Forster evaluated the process thermodynamics of the ered by Warren et al. Additionally, CH4:O2 greater than two results
SnO2/SnO/Sn cycle for both methano-thermal and carbothermal in the formation of solid carbon and thermal losses due to sensible
reduction (Forster, 2004). He reports closed cycle efficiencies for heating of unreacted methane. Further differences between pre-
conversion of solar-to-chemical to electrical energy of 2325% for dicted efficiencies in the two studies arise from significantly higher
the methano-thermal water splitting cycle with SnO2, based on radiation and convection losses assumed by Krenzke and Davidson
Eq. (10). (2014).
In the analysis by Krenzke et al. (2016), efficiency for ceria-
WFC gREACTOR based SCLR is projected for measured conversions and selectivities
g 10
Q SOLAR obtained for a range of methane flow rates through a bed of ceria
fiber particles at 1273 K. For an unoptimized reactor, the highest
The results are based on the assumptions that chemical to elec-
efficiency projected for carbon dioxide splitting is 27% for
trical conversion takes place in a 70% efficient fuel cell and 25%
XCH4 = 25%, XOX = 84%, SH2 = 62%, and SCO = 66%. An expanded anal-
(1000 K) or 35% (1200 K) of the supplied solar energy is lost from
ysis is presented here for an open system with inclusion of the
the reactor due to reflection, radiation, and conduction.
work to extract syngas from the products of (R6). Parameters that
Krenzke et al. (2016) presented a framework for projecting the
are fixed are listed in Table 1. The heat and work requirements are
efficiency, as defined in Eq. (2), of SCLR with respect to conversions
expanded in Eq. (13).
and selectivities (see Eqs. (3)(7)) for an isothermal open system.
The predicted efficiencies account for parasitic work requirements WDP WSEP
to pump gases and to extract pure hydrogen or carbon monoxide. Q SOLAR W=gS!E Q RAD Q LOSS Q GAS Q REJECT
gS!E gS!E
Work to separate syngas from the products of (R6) was not
13
included in the prior work. The framework is transferable to other
metal oxides because the theoretical relationship between effi- On the left hand side, the solar input is that available at the
ciency, conversions, and selectivities is independent of the metal aperture of the reactor. We assume the work input is supplied by
oxide. Krenzke et al. (2016) applied the framework to project the- solar thermal electricity with a solar to electric efficiency (gS?E)
oretical efficiencies over a wide range of conversions and selectiv- of 25%, which is the annual efficiency estimated for dish-Stirling
ities. To interpret the results for a specific material (both systems (Mancini et al., 2003).
composition and morphology), one must know the actual conver- On the right hand side, the heat lost by radiation, QRAD, is esti-
sions and selectivities at specific operating conditions in a specific mated by assuming the receiver approaches blackbody behavior
reactor. and that there are no reflection losses.

Please cite this article in press as: Krenzke, P.T., et al. Solar fuels via chemical-looping reforming. Sol. Energy (2017), http://dx.doi.org/10.1016/j.
solener.2017.05.095
P.T. Krenzke et al. / Solar Energy xxx (2017) xxxxxx 7

Table 1 1
XCO
RT1 XCO2 lnXCO2 1  XCO2 ln1  XCO2
Fixed parameters for projected efficiency of SCLR. WSEP;SPLITTING  2

gSEP
T 1273 K
C 1000 suns 18
I 1000 W m2
The separation of CO2 from CO can be accomplished commer-
P 1 bar
DP 0.05 bar cially with second law efficiencies ranging from 10 to 25%, with
F 0.05 the highest efficiency for amine-based separation processes
Oxidizer H2O (House et al., 2011).
gS!E 25%
The question of post treatment of products of reaction (R6) has
gSEP 10%
gPUMP 80%
not been adequately addressed in the literature. Chemical thermo-
dynamics suggest that high conversion and selectivities are possi-
ble, in which case the question is moot. However, high methane
conversion and syngas selectivities in reaction (R6) have not been
demonstrated together with high conversion of oxidizer in reaction
rT4CAVITY (R7). A possible three-step separation process is depicted in Fig. 9.
Q RAD Q SOLAR 14
IC In the first step, water is condensed from the products. The second
We assume a concentration ratio of 1000, which is achievable separation step removes CO2 from the products. The third removes
with central tower receiver systems. Conduction through the insu- methane from the syngas.
lation and convection losses from the aperture of a solar reactor The work associated with the separation of carbon dioxide from
operating SCLR are estimated as a fraction, F, of the total solar the partial oxidation products can be estimated by
input. n2 RT1 yCO2 lnyCO2 1  yCO2 ln1  yCO2
WSEP;CO2  19
Q LOSS FQ SOLAR 15 gSEP
where the moles of gas entering the second separation step per
Krenzke and Davidson (2014) assumed F = 17% for a laboratory-
mole of oxidizer converted to fuel in the splitting reaction (n2) is
scale prototype reactor. Here we use 5%, which is consistent with
given by Eq. (20) and the carbon dioxide mole fraction at the start
predictions for commercial-scale solar reactors (Schunk et al.,
of the second separation step is given in Eq. (21). The corresponding
2009). The energy requirement associated with mass flows cross-
number of moles of methane that is supplied to (R6) is given in
ing the system boundary, QGAS, is the difference in enthalpy enter-
Eq. (8).
ing and exiting the system.
X X n2 1 2XCH4 SH2 nCH4 ;IN 20
Q GAS ni;OUT hi T  ni;IN hi T1 16
i i XCH4 1  SCO nCH4 ;IN
yCO2 21
This term includes both chemical and sensible energy storage n2
for (R6) and (R7). The heat rejection, QREJECT, is introduced to satisfy In the third separation step, work is expended to remove
an energy balance on the splitting reactor. If the exothermic split- methane from the syngas.
ting reaction releases more energy than is needed to heat the sup-
plied oxidizer to the reaction temperature, heat must be rejected. If n3 RT1 yCH4 lnyCH4 1  yCH4 ln1  yCH4
WSEP;CH4  22
the splitting reaction is carried out at a lower temperature than the gSEP
partial oxidation reaction (R6), it could be necessary to reject sen-
In the absence of reported efficiencies, the efficiency for this
sible heat from the oxide. Two temperature cycling provides favor-
step is assumed to be 10%, which falls within the 540% reported
able chemical thermodynamic conditions for achieving high
by House et al. to typify real-world separation processes
conversions of water to hydrogen in the splitting reaction and
(House et al., 2011). The total moles of gas entering the third sep-
has been considered widely in solar thermochemical metal oxide
aration step (n3 ) and the corresponding methane mole fraction
redox cycles (Lapp et al., 2012; Krenzke and Davidson, 2015;
(yCH4 ) are given by Eqs. (23) and (24).
Ermanoski et al., 2013; Bulfin et al., 2015; Brendelberger et al.,
2015; Lin and Haussener, 2015). Prior thermodynamic process n3 1  yCO2 n2 23
analyses of SCLR have assumed that the released heat remains with
the system and offset the endotherm of (R6), excepting the analy- 1  XCH4 nCH4 ;IN
ses by Krenzke and Davidson (2014) and Warren et al. (2017). To yCH4 24
n3
be consistent with the majority of published work, and because
the magnitude of QREJECT is tied to the specific metal oxide, we Fig. 10 shows the impact of selectivity, with S = SH2 = SCO, and
assume QREJECT = 0. separation work on efficiency as a function of methane conversion.
Work is required for pumping and separation processes. The Solid curves show the efficiency with all separation work included
work required for pumping of gases is given by Eq. (17). for co-production of syngas and hydrogen. Dashed curves show the
  efficiency relevant to operation aimed at producing only hydrogen
P1 DP
nCH4 ;IN nH2 O;IN RT1 ln P1
from water splitting. In that case, one could combust the products
WDP 17 of (R6) without the need for gas separation. For all selectivities,
gP
increasing methane conversion drives increasing efficiency. The
The impact of pressure drop is included in the present analysis ideal SCLR process has an efficiency of 54% at S = XCH4 = 1. The dif-
and found to be negligible relative to the impact of conversions and ference between the solid and dashed curves for S = 1 illustrates
selectivities. the impact of the work required to separate unreacted methane
Separation work is required for both reactions. For water from the synthesis gas. The difference between the solid and
splitting, hydrogen is separated from water by condensation; dashed curves for S = 0.5 shows the combined impact of separating
WSEP,SPLITTING = 0. In the case of CO2 splitting, the required carbon dioxide and methane from the syngas products. For S = 0,
separation work is given by Eq. (18). the separation work accounts for separation of unreacted methane

Please cite this article in press as: Krenzke, P.T., et al. Solar fuels via chemical-looping reforming. Sol. Energy (2017), http://dx.doi.org/10.1016/j.
solener.2017.05.095
8 P.T. Krenzke et al. / Solar Energy xxx (2017) xxxxxx

H2,CO,H2O,CO2,CH4 H2,CO, CO2,CH4 H2,CO, CH4 H2,CO

H2O CO2 CH4

Separaon 1 Separaon 2 Separaon 3

Fig. 9. Conceptual process for extraction of syngas from the products of partial oxidation of methane.

0.6 0.6
S=1
XH2O=1
0.5 0.5

0.4 0.5
S=0.5 0.4

0.25

0.3


0.3

S=0
0.2 0.2 0.1
Includes
0.1 Separation
Work 0.1

0 0
0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
X CH X CH
4
4

Fig. 10. Efficiency as a function of methane conversion for cycling at 1273 K and Fig. 11. Efficiency as a function of methane conversion for cycling at 1273 K with
XH2 O = 1, S SH2 = SCO. (Other parameters are listed in Table 1.) Solid curves show S = 1. Curves are labeled with the oxidizer conversion (XH2 O ). Other parameters are
the efficiency with all separation work included for co-production of syngas and listed in Table 1.
hydrogen. Dashed curves show the efficiency relevant to operation aimed at only
splitting water.

and products, i.e. for both (R6) and (R7), per unit mass of the metal
from carbon dioxide. Notably, the projected efficiencies for S = 0.5 oxide oxygen carrier. The reactor mass is proportional to efficiency,
and S = 0 are close when the separation work is included. Improve- but decreases as the rates of syngas production increase. Thus,
ments in selectivity reduce the sensible heating requirement for improving reaction rates has a positive and direct impact on the
SCLR because a larger fraction of the product gases possesses heat- practicality of implementing SCLR in a solar reactor. A preliminary
ing value. However, as selectivity is increased from 0 to 0.5, estimate of the mass of reactive material in the reactor is found for
increased separation work offsets the reduction in sensible the ideal efficiency eta = 54%. Assuming simultaneous operation of
heating. the partial oxidation and splitting reactions at 1273 K and a sup-
The dependence of efficiency on oxidizer conversion in the plied methane flow rate of 10 mL min1 g1, the estimated mass
water splitting reaction is shown in Fig. 11 for S = SH2 = SCO = 1. is 64 kg per megawatt of solar thermal input.
The present thermodynamic analysis indicates SCLR has the
The dramatic increase in efficiency with increasing oxidizer con-
potential for high solar-to-chemical efficiency. The upper limit is
version is due to the decrease in sensible and latent heating load.
approximately 54% for reasonable assumptions of operating condi-
For XCH4 = 1, the efficiency increases from 18% to 54% as oxidizer
tions as listed in Table 1. Key performance parameters are the
conversion increases from XH2 O = 0.1 to 1.
methane conversion (XCH4 ), hydrogen selectivity (SH2 ), and carbon
Looking toward the design of solar reactors for SCLR, we esti-
monoxide selectivity (SCO ) for the partial oxidation reaction (R6)
mate the mass of metal oxide required. From the solar-to-
and the oxidizer conversion (XH2 O ) for the splitting reaction (R7).
chemical efficiency of Eq. (2), the estimated reactor mass per
  Efficiency is most sensitive to the methane and oxidizer conver-
megawatt of total solar thermal input Q_ SOLAR W is given by
_
gS!E sions. The conversion of oxidizer in (R7) is negatively impacted
Eq. (25). by CH4:O2 ratios below 2:1 in (R6).


m
P 0
g
25
_
Q SOLAR gW_
i n_ i;OUT  n_ 0i;IN HHVi
S!E
3. Oxygen carrier materials

In Eq. (25), the solar thermal input (Q_ SOLAR ) and solar thermal Prior reviews of oxygen carrier materials for solar thermochem-
equivalent of work requirements (W= _ g ) are in rate form. The ical metal oxide redox cycles for water and carbon dioxide (Scheffe
S!E
terms n_ 0i;IN and n_ 0i;OUT represent the molar flow rates of reactants and Steinfeld, 2014; Roeb et al., 2012; Agrafiotis et al., 2015; Miller

Please cite this article in press as: Krenzke, P.T., et al. Solar fuels via chemical-looping reforming. Sol. Energy (2017), http://dx.doi.org/10.1016/j.
solener.2017.05.095
P.T. Krenzke et al. / Solar Energy xxx (2017) xxxxxx 9

et al., 2014; Muhich et al., 2016) and CLR (Li et al., 2013a; Protasova (In2O3/In, SnO2/Sn) products or solid products with an altered crys-
and Snijkers, 2016; Adanez et al., 2012; Voitic and Hacker, 2016) tal structure (Fe3O4/FeO). Non-volatile metal oxides can sinter at
emphasize the importance of the metal oxide for efficient opera- the temperatures anticipated for CLR and are often supported on
tion of these cycle. This emphasis is equally valid for SCLR. inert structures (e.g. SiO2, Al2O3, ZrO2). Methano-thermal reduction
Methane conversion, H2 and CO selectivities, and, to a lesser extent of some metal oxides forms metal carbides, for which oxidation is
oxidizer conversion, carbon deposition and morphological stability less favored thermodynamically than the corresponding reduced
have been evaluated for a number of material, including stoichio- metal oxide (e.g. V8C7, WC, and Fe3C which form from V2O5,
metric metal oxides (Steinfeld et al., 1996, 1995a, 1995b, 1993, WO3, and Fe3O4, respectively) (Kodama, 2000; Steinfeld and
1998; Forster, 2004; Steinfeld, 1997; Go et al., 2009; Steinfeld Meier, 2004).
and Spiewak, 1998; Krupl and Steinfeld, 2001a, 2001b; Kodama, Solar CLR was first proposed by Steinfeld et al. with Fe3O4
2000; Shimizu et al., 2001; Kodama et al., 2003; Sim et al., 2010; (Steinfeld, 1997; Steinfeld et al., 1995a, 1993). Fe2O3 reduces/oxi-
Jang et al., 2015; Stobbe et al., 1999; de Diego et al., 2008), ferrites dizes in a stepwise fashion (Fe2O3 M Fe3O4 M FeO M Fe). The
(Kodama et al., 2002; Takenaka et al., 2004, 2005; Kang et al., 2008, reduction from Fe2O3 to Fe3O4 favors complete combustion prod-
2010; Cha et al., 2009, 2010; Scheffe et al., 2011; Go et al., 2008), ucts. Complete reduction to metallic Fe often produces solid car-
mixed metal oxides (Nakayama et al., 2010; Dueso et al., 2012; bon. On the other hand, reoxidation to Fe2O3 is slow. Therefore,
Rydn et al., 2008; Suzuki et al., 2012), ceria-based oxides Fe3O4/FeO is the preferred redox pair (Go et al., 2009). It yields
(Otsuka et al., 1993, 1998, 1999; Krenzke et al., 2016; Warren low syngas selectivity and low water conversion (<30%)
et al., 2017; Sang et al., 2014; Sadykov et al., 2004; Wei et al., (Steinfeld et al., 1995a, 1993; Go et al., 2009; Suzuki et al., 2012).
2007, 2010; Li et al., 2008a, 2008b, 2010a, 2010b, 2011a, 2013b; Steinfeld et al. also considered ZnO for SCLR (Steinfeld et al.,
Gu et al., 2013, 2014, 2016; Kim et al., 2010; Jha et al., 2015; Zhu 1995b, 1996, 1998; Steinfeld, 1997; Steinfeld and Spiewak, 1998;
et al., 2010, 2011a, 2011b, 2014, 2013; Nair and Abanades, 2016; Krupl and Steinfeld, 2001a, 2001b). Methane conversion was
Gonzlez-Velasco et al., 1999; Jiang et al., 2014a, 2008; Abanades 25% at 1273 K and nearly 100% at 1573 K (Steinfeld et al.,
and Flamant, 2006a, 2006b; He et al., 2009, 2013a, 2013b; Jeong 1995b). Carbon deposition was observed but not quantified. The
et al., 2011; Zheng et al., 2014a, 2014b; Fathi et al., 2000; Gao methano-thermal reduction step was implemented in a vortex
et al., 2016; Pantu et al., 2000), and perovskites (Rydn et al., flow solar reactor (Krupl and Steinfeld, 2001a, 2001b; Steinfeld
2008; Nalbandian et al., 2009, 2011; Evdou et al., 2008, 2010; Li et al., 1998). Steinfeld et al. considered Zn(s) more valuable than
et al., 2002, 2004; Dai et al., 2008a, 2008b, 2008c, 2016, 2006a, syngas and thus the reactor was operated with CH4:ZnO > 10:1.
2006b; Wei et al., 2008; Zhao et al., 2014, 2016; Mihai et al., Zn yields were 83100%. Methane conversion was low due to the
2012; Mudu et al., 2016; Mishra et al., 2016; Zeng et al., 2003; higher than stoichiometric ratio of CH4:ZnO.
Garca et al., 2010; Shafiefarhood et al., 2015; Franca et al., 2012; Kodama et al. compared the performance of Fe3O4, ZnO, In2O3,
Michalsky et al., 2015). The studies were conducted over a wide SnO2, V2O5, MoO2, and WO3 for methane partial oxidation at
range of operating conditions. As a consequence, direct comparison 1173 K (Kodama, 2000). WO3 provided the best combination of
of materials is not always meaningful. conversion and selectivities for the reduction step and suitability
Prior experimental studies for which conversions and selectivi- for water splitting. Supporting WO3 on ZrO2 improved methane
ties are reported or could be extracted from the raw data are listed conversion (from 40% to 70%), oxidizer conversion (from 7% to
in Tables 29. Table 2 lists the studies for stoichiometric metal oxi- 30%), and H2 selectivity (from 69% to 97%) and decreased CO selec-
des. Table 3 lists the studies for ferrites and mixed metal oxides. tivity (from 97% to 86%). The improvements were attributed, in
Tables 47 list the studies for pure, catalyst promoted, and doped part, to the interaction between WO3 and the inert support. Tung-
ceria. Table 8 lists the studies for perovskites with Fe B-site cations sten carbide (WC) was formed when WO3/ZrO2 was reduced above
(AFeO3). Table 9 lists the studies for perovskites with Mn B-site 1123 K (Shimizu et al., 2001). In experiments at 1350 K in a solar
cations (AMnO3). The material morphology (particle size and speci- simulator, methane conversions up to 93% and H2 and CO selectiv-
fic surface area) and operating conditions for both reaction steps ities of 46% and 71% are reported for 50 wt.% WO3/ZrO2 (Kodama
(temperature, reaction time, reactant concentration, and total flow et al., 2003). Solar-to-chemical efficiencies would be lowered how-
rate) are provided when available. The redox performance is quan- ever by poor water splitting performance; H2O conversion was
tified by the methane and oxidizer conversions, and selectivities <30%.
towards hydrogen, carbon monoxide, and carbon. If values were Doping Fe3O4 improves redox performance and stability
not reported, we present values calculated from the published data (Kodama et al., 2002; Takenaka et al., 2004, 2005; Kang et al.,
(graphical data were digitized to extract numerical values). When 2008, 2010; Cha et al., 2009, 2010; Scheffe et al., 2011; Go et al.,
necessary, material balances were used to quantify unreported 2008). Kodama et al. (2002) evaluated M0.39Fe2.61O4 for M = Ni,
species. For some entries in the tables, reported values differ from Zn, Co. The Ni-ferrite yielded the highest conversion and selectivi-
those reported in the literature due to differences in parameter ties (XCH4 = 31%, SH2 = 75%, SCO = 72%). Support on ZrO2 suppressed
definitions. Unless noted, all data presented in the tables are con- sintering and improved reactivity. In five successive CLR cycles
sistent with the definitions provided in the present review. with reduction at 1173 K and oxidation at 1073 K, carbon deposi-
tion led to partial deactivation. At 1350 K in a solar simulator,
3.1. Stoichiometric materials methane conversions were as high as 94% and syngas selectivity
was 60%. Cha et al. (2009) considered Cu-ferrites. Methane conver-
Stoichiometric metal oxides were the first oxygen carrier mate- sion increased with increasing Cu content and syngas selectivity
rials considered for CLR (Steinfeld et al., 1996, 1995a, 1995b, 1993, was highest for Cu0.7Fe2.3O4/ZrO2. In comparison to Fe3O4/ZrO2,
1998; Forster, 2004; Steinfeld, 1997; Go et al., 2009; Steinfeld and CO selectivity was higher due to less carbon deposition. Methane
Spiewak, 1998; Krupl and Steinfeld, 2001a, 2001b; Kodama, 2000; and hydrogen selectivities were lower.
Shimizu et al., 2001; Kodama et al., 2003; Sim et al., 2010; Jang Rydn et al. reports mixed metal oxides of NiO, Fe2O3, and
et al., 2015; Stobbe et al., 1999; de Diego et al., 2008; Suzuki Mn3O4 yield low syngas selectivity (Rydn et al., 2008). Suzuki
et al., 2012). Because they can be fully reduced to their metallic et al. found NiO-based oxides yield higher methane conversion
components, stoichiometric metal oxides generally have high and syngas selectivity than Fe2O3-based mixed oxides. Ni-Cr-
oxygen storage capacity. Upon reduction, stoichiometric metal MgO oxides may be promising if carbon deposition can be avoided
oxides undergo phase change to either gaseous (ZnO/Zn) or liquid (Nakayama et al., 2010; Suzuki et al., 2012).

Please cite this article in press as: Krenzke, P.T., et al. Solar fuels via chemical-looping reforming. Sol. Energy (2017), http://dx.doi.org/10.1016/j.
solener.2017.05.095
Table 2

10
solener.2017.05.095
Please cite this article in press as: Krenzke, P.T., et al. Solar fuels via chemical-looping reforming. Sol. Energy (2017), http://dx.doi.org/10.1016/j.

Stoichiometric metal oxides.

Material Reduction Oxidation Ref.


Composition Size SSA yCH4 V_ t T XCH4 SH2 SCO Sc OX yox V_ t T Xox
 2    
(m) m (mol%) ml (min) (K) (%) (%) (%) (%) (mol%) ml (min) (K) (%)
g ming ming

Fe3O4 2 10 100 9 1073 20 41 42 0.00 NA Steinfeld et al. (1993)


Fe3O4 2 2.31 4333 50 1273 16 16 18 69 NA Steinfeld et al. (1995a)
1373 32 48 52 45
1473 82 58 29 65
1573 97 57 24 67
Fe2O3 20 50 10 973 21 07 10 NA Suzuki et al. (2012)
3:7 Fe2O3/Al2O3 100 10 12 1123 68 87 42 47 NA Li et al. (2010a)
3:7 Fe2O3/Al2O3 100 10 17 1123 48 72 45 30 NA Li et al. (2010b)
ZnO 1.2 5 4333 10 1273 25 NA Steinfeld et al. (1995b)
1373 51
1473 84
1573 100

P.T. Krenzke et al. / Solar Energy xxx (2017) xxxxxx


Fe3O4 50 12 60 1173 15 58 63 NA Kodama (2000)
ZnO 9 07 00 91
SnO2 18 22 25 28
In2O3 8 14 65 62
MoO2 11 16 100 83
V2O5 19 31 100 84
WO3 9 19 55 92
WO3 50 12 60 1273 40 69 97 H2O 47 23.4 160 1073 7 Kodama (2000)
50 wt.% WO3/SiO2 6 44 61 87 11.7 11
50 wt.% WO3/Al2O3 6 49 66 84 11.7 13
50 wt.% WO3/ZrO2 6 70 97 86 11.7 30
50 wt.% WO3/ZrO2 50 6 60 1073 10 54 74 0 NA Kodama (2000)
1123 52 76 50 45
1173 57 61 50 45
1223 68 71 44 49
1273 70 66 51 24
50 wt.% WO3/ZrO2 50 4 48 1173 21 77 78 3 H2O 47 4 120 1173 7 Kodama et al. (2003)
24 100 75 8 8
27 94 69 15 10
30 86 80 6 8
29 100 73 11 12
50 4 48 1273 56 75 83 0 H2O 47 4 140 1273 18
48 87 79 0 12
49 75 81 5 10
50 80 86 3 15
42 87 80 6 16
50 wt.% WO3/ZrO2 50 6 6 1335 58 50 79 1 NA Kodama et al. (2003)
12 3 1359 81 46 74 0
24 1.5 1382 85 47 77 1
48 0.75 1350 93 46 71 1
Table 3
solener.2017.05.095
Please cite this article in press as: Krenzke, P.T., et al. Solar fuels via chemical-looping reforming. Sol. Energy (2017), http://dx.doi.org/10.1016/j.

Ferrites and mixed metal oxides.

Material Reduction Oxidation Ref.


Composition Size SSA yCH4 V_ t T XCH4 SH2 SCO Sc OX yox V_ t T Xox
 2    
(mm) m (mol%) ml (min) (K) (%) (%) (%) (%) (mol%) ml (min) (K) (%)
g ming ming

Co0.39Fe2.61O4 < 0.05 52 12 60 1173 17 51 52 NA Kodama et al. (2002)

Zn0.39Fe2.61O4 6 55 43
Ni0.39Fe2.61O4 31 75 72
33 wt.% Ni0.39Fe2.61O4/ZrO2 52 4 48 1173 58 46 44 25 H2O 47 7.8 120 1073 22 Kodama et al. (2002)
47 47 48 24 18
47 83 45 34 21
21
46 48 47 30 19
53 68 47 30
33 wt.% Ni0.39Fe2.61O4/ZrO2 52 8 6 9851001 56 55 34 11 NA Kodama et al. (2002)
10331072 86 60 53 1
11301188 97 54 52 1
11501235 96 47 44 1

P.T. Krenzke et al. / Solar Energy xxx (2017) xxxxxx


12641302 100 30 36 2
Ni0.39Fe2.61O4/ZrO2 52 64 0.75 12581322 94 24 28 1 NA Kodama et al. (2002)
12461374 94 54 54 1
12971400 93 60 60 1
CuFe2O4 0.42 51.6 10 50 1173 56 48 55 4 H2O 20 12.5 120 1073 Kang et al. (2008)
34 50 57 17
37 75 55 24
46 52 60 15
55 53 59 14
60 wt.% Fe3O4/ZrO2 250425 15 30 120 1173 50 83 33 52 H2O 31 43.5 120 973 Cha et al. (2009)
60 wt.% Cu0.3Fe2.7O4/ZrO2 12 26 36 0
60 wt.% Cu0.5Fe2.5O4/ZrO2 12 30 39 0
60 wt.% Cu0.7Fe2.3O4/ZrO2 18 54 61 0
60 wt.% Cu01.0Fe2.0O4/ZrO2 25 64 51 12
60 wt.% Cu1.5Fe1.5O4/ZrO2 38 72 27 47
60 wt.% Cu0.7Fe2.3O4/ZrO2 250425 15 30 120 1173 18 54 61 0 H2O 31 43.5 120 973 Cha et al. (2009)
60 wt.% Cu0.7Fe2.3O4/Ce0.5Zr0.5O2 21 49 49 10
60 wt.% Cu0.7Fe2.3O4/Ce0.67Zr0.33O2 68 82 22 61
60 wt.% Cu0.7Fe2.3O4/Ce0.75Zr0.25O2 64 80 28 56
60 wt.% Cu0.7Fe2.3O4/CeO2 6 13 25 27
20 wt.% CuFe2O4/ZrO2 51.6 10 9 1173 82 58 48 37 H2O 20 22.5 80 1073 Kang et al. (2010)
83 69 48 42
80 68 47 43
77 64 45 46
74 65 46 42
20 wt.% CuFe2O4/CeO2 51.6 10 9 1173 57 38 44 38 H2O 20 22.5 80 1073 Kang et al. (2010)
60 55 65 28
78 80 78 17
84 83 81 15
88 83 81 14
89 81 79 16
89 85 83 12
91 82 80 16
92 82 81 15

(continued on next page)

11
12
Table 3 (continued)
solener.2017.05.095
Please cite this article in press as: Krenzke, P.T., et al. Solar fuels via chemical-looping reforming. Sol. Energy (2017), http://dx.doi.org/10.1016/j.

Material Reduction Oxidation Ref.


Composition Size SSA yCH4 V_ t T XCH4 SH2 SCO Sc OX yox V_ t T Xox
 2    
(mm) m (mol%) ml (min) (K) (%) (%) (%) (%) (mol%) ml (min) (K) (%)
g ming ming

90 88 87 9
90 84 82 14
Fe2O3-NiO (2:1) 20 100 10 1073 0.59 0.06 0.10 O2 20 100 10 1073 NA Suzuki et al. (2012)
Fe2O3-Co3O4 (2:1) 0.25 0.06 0.09
Fe2O3-Cr2O3 (2:1) 0.21 0.25 0.20
Fe2O3-MgO-Cr2O3 (3:1:2) 0.45 0.73 0.63
Fe2O3-CaO-Cr2O3 (3:1:2) 0.29 0.09 0.11
Fe2O3-SrO-Cr2O3 (3:1:2) 0.24 0.08 0.09
NiO-MgO (16:25) 20 50 10 973 0.78 0.98 0.30 0.65 O2 20 50 10 973 NA Nakayama et al. (2010)
NiO-MgO (16:25) 0.98 0.96 0.34 0.60 and Suzuki et al. (2012)
NiO-Al2O3-MgO (16:4:25) 0.76 0.98 0.47 0.51
NiO-CaO-MgO (16:4:25) 0.84 0.97 0.38 0.59
NiO-Fe2O3-MgO (16:4:25) 0.55 0.57 0.34 0.24
NiO-Co3O4-MgO (16:4:25) 0.74 0.98 0.28 0.69
NiO-Cr2O3-MgO (16:4:25) 0.98 0.71 0.43 0.20

P.T. Krenzke et al. / Solar Energy xxx (2017) xxxxxx


NiO-Cr2O3-MgO (16:4:25) 0.93 0.75 0.64 0.11
Ni-Cr2O3-SiO2 (16:4:25) 0.94 0.71 0.52 0.07
Ni-Cr2O3-CeO2 (16:4:25) 0.92 0.48 0.40 0.11
Ni-Cr2O3-Y2O3 (16:4:25) 0.83 0.69 0.61 0.04
Ni-Cr2O3-La2O3 (16:4:25) 0.92 0.50 0.29 0.26
Ni-Cr2O3 (16:4) 0.93 0.28 0.11 0.18
52.6% NiO- 6.6% Cr2O3-MgO 20 50 10 873 40 94 58 32 O2 20 50 10 873 NA Nakayama et al. (2010)
923 77 92 67 19 923 and Suzuki et al. (2012)
973 93 82 66 10 973
1023 97 70 57 10 1023
1073 98 61 47 14 1073
solener.2017.05.095
Please cite this article in press as: Krenzke, P.T., et al. Solar fuels via chemical-looping reforming. Sol. Energy (2017), http://dx.doi.org/10.1016/j.

Table 4
Pure CeO2.

Material Reduction Oxidation Ref.


Composition Size SSA yCH4 V_ t T XCH4 SH2 SCO Sc OX yox V_ t T Xox
 2  
(mm) m (mol%) ml (min) (K) (%) (%) (%) (%) (mol%) ml (min) (K) (%)
g ming ming

P.T. Krenzke et al. / Solar Energy xxx (2017) xxxxxx


CeO2 0.40.9 0.82 50 120 160 973 1 98 94 CO2 12.5 40 60 723 ca. 100 for 10 min Otsuka et al. (1993,
H2O 2.4 40 80 823 ca. 1.00 for 40 min 1997, 1998)
CeO2 100 10 12 1123 24 87 92 0 NA Li et al. (2010a)
1 1
CeO2 0.40.9 21 100 10 9 1123 51 87 87 6 0.1 ml min g liquid 30 973 Zhu et al. (2011b)
water in 28 ml min1 g1 N2
CeO2 0.40.9 100 10 20 1123 31 96 73 20 0.1 ml min1 g1 liquid 30 1123 Zhu et al. (2013)
water in 28 ml min1 g1 N2
CeO2 5 (6 m fibers) 0.143 100 1 8 1173 20 36 37 0 CO2 100 10 25 1173 31 Krenzke et al. (2016)
2.5 11 47 47 0 40
5 7 56 53 0 50
10 4 68 59 0 57
15 2 78 59 0 49
1 1273 60 40 39 0 1273 48
2.5 37 56 54 0 56
5 25 62 66 0 84
10 17 71 74 0 90
15 13 77 78 0 93
CeO2 0.41.0 10 4.4 120 1223 9 64 58 8 O2 100 0.4 5 1223 Warren et al. (2017)
1308 41 85 75 8 O2 100 0.4 20 1308
1393 52 83 59 26 CO2 100 0.44.4 120 1393 21
20 wt.% CeO2/c-Al2O3 158 100 0.5 ml g1 pulses 28 pulses 923 14 17 30 NA Fathi et al. (2000)

13
14
solener.2017.05.095
Please cite this article in press as: Krenzke, P.T., et al. Solar fuels via chemical-looping reforming. Sol. Energy (2017), http://dx.doi.org/10.1016/j.

Table 5
Catalyst promoted CeO2.

Material Reduction Oxidation Ref.


Composition Size SSA yCH4 V_ t T XCH4 SH2 SCO Sc OX yox V_ t T Xox
 2    
(mm) m (mol%) ml (min) (K) (%) (%) (%) (%) (mol%) ml (min) (K) (%)
g ming ming

P.T. Krenzke et al. / Solar Energy xxx (2017) xxxxxx


20 wt.% CeO2/c-Al2O3 158 100 0.5 ml g1 pulses 28 pulses 923 14 17 30 NA Fathi et al. (2000)
1
0.5 wt.% Rh/CeO2/c-Al2O3 157 100 0.5 ml g pulses 28 pulses 923 95 35 51 NA Fathi et al. (2000)
0.5 wt.% Pt- CeO2/c-Al2O3 158 100 0.5 ml g1 pulses 28 pulses 923 95 92 56 27 CO2 100 0.5 ml g1 pulses 5 873 100 Fathi et al. (2000)
10 90
15 67
20 54
25 45
5 973 100
10 100
15 83
20 67
25 55
0.5 wt.% Pt/CeO2 0.30.7 40.4 5 500 7 873 20 96 88 6 O2 5 500 873 NA Pantu et al. (2000)
14 96 91 3
20 4 98 88 7
6 98 87 7
5 973 22 97 89 6 973
21 97 93 3
20 4 98 84 10
4 97 83 11
0.5 wt.% Pt/CeO2 0.30.7 40.4 100 5 ml g1 pulses 30 pulses 823 10 95 82 9 O2 100 5 ml g1 pulses 823 NA Pantu et al. (2000)
P.T. Krenzke et al. / Solar Energy xxx (2017) xxxxxx 15

There are limited data for water and carbon dioxide splitting

Zheng et al. (2014a, 2014b)

Zheng et al. (2014a, 2014b)


with ferrites and mixed metal oxides following reduction with
methane. The highest demonstrated oxidizer conversion is 22%
for water splitting with 33 wt.% Ni0.39Fe2.61O4/ZrO2 (Kodama

Pantu et al. (2000)


Zhu et al. (2011a)
et al., 2002). As with WO3/ZrO2, low oxidizer conversion is likely

Li et al. (2010b)
to limit the potential for achieving high efficiencies with ferrites
and mixed metal oxides.
Ref.

3.2. Nonstoichiometric materials


Xox

NA
(%)

In contrast to stoichiometric metal oxides, nonstoichiometric


metal oxides such as ceria and perovskites do not undergo struc-


1073

1073
973

873
tural phase transitions between the fully oxidized and partially
(K)
T

reduced forms over a wide nonstoichiometry (d) range. On the


(min)

other hand, nonstoichiometric oxygen carriers generally exhibit


45

45
30

lower oxygen storage capacity. Although oxygen diffusion through



t

the lattice is fast, prior study of catalyzed and uncatalyzed materi-


10 ml g1

als indicates reduction by methane can be kinetically limited


pulses

ming

(Otsuka et al., 1998; Fathi et al., 2000; Pantu et al., 2000). This
ml

V_

observation is supported by data showing increasing rates of syn-


gas production with increased surface area (Otsuka et al., 1998; Li
(mol%)

et al., 2011b; Zheng et al., 2014b; Gao et al., 2016; Pantu et al.,
100
yox

2000; Evdou et al., 2010; Dai et al., 2008c).


Reduction is suggested to proceed via a modified Mars-van
0.1 ml min g liquid water

0.1 ml min1 g1 liquid water

0.1 ml min1 g1 liquid water

Krevelen mechanism, with reaction steps given by (R8)(R13)


(Otsuka et al., 1998; Pantu et al., 2000; Dai et al., 2006a).
in 67 ml min1 g1 N2

in 67 ml min1 g1 N2

CH4 5 ! CH3 H 3 !    ! C 4H R8


in 28 ml min1 N2
1
1

C O surface ! CO  ! COg 2 R9


Oxidation

C 2O surface ! CO2 2 ! CO2 g 3 R10


NA
OX

O2

2H ! H2  ! H2 g 2 R11


(%)

34
Sc

6
9
0

2H O surface ! H2 O 2 ! H2 Og 3 R12


0.73
SCO
(%)

87

91

39

77

73

81
81
80
80

70
90

Obulk  ! O surface R13


0.85
(%)
SH2

87

67

89

89
92
94
95

81
81
80

90

At the initiation of reduction, surface adsorbed oxygen is abundant


0.42
XCH4

and chemisorbed carbon and hydrogen are readily oxidized to CO2


(%)

51

41

56
54
52

36
60

40

8
7

and H2O. Consequentially, methane conversions are high but syngas


1123

1073

1073
1073

selectivities are low. As the reaction proceeds, surface adsorbed


873
(K)
T

oxygen is consumed faster than it can be replenished by the diffu-


sion of bulk lattice oxygen to the surface. The lower availability of
pulses
(min)

surface oxygen leads to lower methane conversion, but higher syn-


17
40

40

40
9

6
t

gas selectivity. When the rate of diffusion of the bulk lattice oxygen
10 ml g1

to the surface is slower than methane dissociation, chemisorbed


pulses


carbon accumulates on the surface. Carbon accumulation tends to


ming
ml

5.6

5.6

occur when more than 80% of the available oxygen from the oxygen
33

33

V_
Reduction

carrier material has been consumed (Otsuka et al., 1993, 1997,


(mol%)

1998; Cho et al., 2005). Transition and platinum-group metals can


yCH4

100

100

100
10

10

catalyze methane decomposition and expedite the onset of carbon


deposition (Fathi et al., 2000). Similarly, the proposed mechanism
Temp

1123
1023
 2

Calc.

for the oxidizer splitting step is given by (R14)(R16), expressed


723
823
923
SSA

g
m

21

85

59
42
40

40

here for water splitting.


0.40.9

0.30.7

H2 Og 2 ! H2 O  ! H2 O surface R14


(mm)
Size

H2 ! H2 g  R15
0.5 wt.% Pt/Ce0.8Zr0.2O2
0.5 wt.% Pt/Ce0.5Zr0.5O2

O surface ! Obulk R16


0.5 wt.% Pt/CeO2
Zr4+ doped CeO2.

Oxidizer conversion is highest at the initiation of the oxidizer


Composition

Ce0.7Zr0.3O2
Ce0.8Zr0.2O2

Ce0.7Zr0.3O2
Ce0.8Zr0.2O2

splitting step when the surface and bulk oxygen in the material
Material

are depleted and decreases as the oxygen vacancies in the material


CeO2
Table 6

lattice are filled and surface and bulk oxygen become more
abundant.

Please cite this article in press as: Krenzke, P.T., et al. Solar fuels via chemical-looping reforming. Sol. Energy (2017), http://dx.doi.org/10.1016/j.
solener.2017.05.095
Table 7

16
solener.2017.05.095
Please cite this article in press as: Krenzke, P.T., et al. Solar fuels via chemical-looping reforming. Sol. Energy (2017), http://dx.doi.org/10.1016/j.

Fe3+ doped CeO2.

Material Reduction Oxidation Ref.


Composition Size SSA yCH4 V_ t T XCH4 SH2 SCO Sc OX yox V_ t T Xox
 2    
(mm) m (mol%) ml (min) (K) (%) (%) (%) (%) (mol%) ml (min) (K) (%)
g ming ming

Ce0.5Fe0.5O2 100 17 11 1093 32 68 76 0 NA Li et al. (2008b)


1123 43 68 80 0
1153 59 88 89 0
Ce0.5Fe0.5O2 100 17 11 1153 58 86 88 0 O2 21 1153 NA Li et al. (2008b)
55 96 97 0
53 99 99 0
CeO2 100 12 12 1123 24 87 92 0 NA Li et al. (2010a)
Ce0.9Fe0.1O2 39 81 34 62
Ce0.8Fe0.2O2 53 74 70 19
Ce0.7Fe0.3O2 54 71 73 14
Ce0.6Fe0.3O2 63 69 53 38
Ce0.5Fe0.5O2 71 66 43 45
Ce0.7Fe0.3O2 100 10 12 1123 NA Li et al. (2010a)

P.T. Krenzke et al. / Solar Energy xxx (2017) xxxxxx


Co-precipitation 54 71 73 14
Hydrothermal 89 73 70 2
Solid Phase 29 86 81 6
Ce0.7Fe0.3O2 100 10 10 1123 89 70 69 0 O2 21 33 16 1123 NA Li et al. (2010a)
88 81 87
87 83 86
83 86 88
83 86 86
83 86 88
Ce0.7Fe0.3O3 100 5.6 17 1073 42 83 84 0 NA Li et al. (2010b)
Ce0.7Fe0.3O3 100 5.6 13 1073 59 88 88 0 O2 21 17 1073 NA Li et al. (2010b)
57 87 89 0
Ce0.7Fe0.3O3 0.40.9 100 5.6 11 1123 51 69 70 0 0.1 ml min1 g1 liquid water 30 min 1073 Zhu et al. (2010)
in 28 ml min1 g1 N2
Ce0.7Fe0.3O3 45 100 5.6 16 1123 59 71 80 1 NA Li et al. (2011b)
26 61 80 83 1
13 63 82 85 3
5 64 88 78 12
Ce0.7Fe0.3O3 13 100 10 13 1123 63 82 87 0 O2 21 60 1123 NA Li et al. (2011b)
64 87 89 0
66 91 92 0
62 92 94 0
62 93 94 0

CeO2 0.40.9 27 100 5.6 20 1123 31 96 73 20 0.1 ml min1 g1 liquid water 30 1123 Zhu et al. (2013)
in 28 ml min1 g1 N2
Ce0.9Fe0.1O2 5 14 88 83 1
Ce0.8Fe0.1O2 11 32 98 88 0
Ce0.7Fe0.3O2 22 41 88 83 1
Ce0.6Fe0.4O2 16 58 82 79 7
Ce0.5Fe0.5O2 15 63 83 83 1
Ce0.4Fe0.6O2 13 71 91 84 1
FeO2 7 21 13 16 0
Ce0.5Fe0.5O2 0.40.9 15 100 5.6 20 1073 39 73 76 0 0.1 ml min1 g1 liquid water 30 1073 Zhu et al. (2013)
in 28 ml min1 g1 N2
Table 7 (continued)
solener.2017.05.095
Please cite this article in press as: Krenzke, P.T., et al. Solar fuels via chemical-looping reforming. Sol. Energy (2017), http://dx.doi.org/10.1016/j.

Material Reduction Oxidation Ref.


Composition Size SSA yCH4 V_ t T XCH4 SH2 SCO Sc OX yox V_ t T Xox
 2    
(mm) m (mol%) ml (min) (K) (%) (%) (%) (%) (mol%) ml (min) (K) (%)
g ming ming

1123 63 83 84 0 1123
1173 77 87 86 0 1173
Ce0.5Fe0.5O2 0.40.9 15 100 5.6 20 1123 39 83 84 0 0.1 ml min1 g1 liquid water 30 1123 Zhu et al. (2013)
in 28 ml min1 g1 N2
30 63 94 80 0
40 77 100 66 20
Ce0.5Fe0.5O2 0.40.9 15 100 5.6 20 1123 63 83 84 0 0.1 ml min1 g1 liquid water 30 1123 Zhu et al. (2014, 2013)
in 28 ml min1 g1 N2
63 88 88
59 88 88
53 87 87
55 90 89
53 87 88
54 89 87
51 88 89

P.T. Krenzke et al. / Solar Energy xxx (2017) xxxxxx


50 86 89
50 86 87
Ce0.7Fe0.3O2 13 100 5.6 18 1073 57 82 84 0 O2 21 33 20 1073 NA Li et al. (2011a)
(Ce0.7Fe0.3)0.99(Zr)0.01O2 13 62 82 83 0
(Ce0.7Fe0.3)0.95(Zr)0.05O2 15 64 84 84 0
(Ce0.7Fe0.3)0.90(Zr)0.10O2 20 64 80 81 0
(Ce0.7Fe0.3)0.95(Zr)0.05O2 15 100 5.6 18 1073 64 84 84 0 O2 21 33 20 1073 NA Li et al. (2011a)
59 89 89 0
58 88 89 0
59 89 90 0
60 90 90 0

17
Table 8

18
solener.2017.05.095
Please cite this article in press as: Krenzke, P.T., et al. Solar fuels via chemical-looping reforming. Sol. Energy (2017), http://dx.doi.org/10.1016/j.

AFeO3 perovskites.

Material Reduction Oxidation Ref.


Composition Size SSA yCH4 V_ t T XCH4 SH2 SCO Sc OX yox V_ t T Xox
 2    
(mm) m (mol%) ml (min) (K) (%) (%) (%) (%) (mol%) ml (min) (K) (%)
g ming ming

LaFeO3 10 100 1 ml g1 pulse 15 pulses 1073 10 61 72 0 NA Dai et al. (2006a)


1123 22 73 87 0
1173 52 89 93 0
LaFeO3 6 11 92 0.5 1173 67 99 0 O2 11 92 0.17 1173 NA Dai et al. (2016)
LaFeO3 10 153 3 1083 23 92 96 O2 10 153 4 1083 NA Li et al. (2004)
1123 60 97 99 1123
1173 59 96 98 1173
LaFeO3 10 153 3 1173 72 96 98 O2 10 153 4 1173 NA Li et al. (2004)
La0.9Sr0.1FeO3 56 91 94
La0.8Sr0.2FeO3 59 91 94
La0.5Sr0.5FeO3 31 84 86
La0.8Sr0.2FeO3 11 153 0.4 1173 90 93 93 0 O2 11 153 0.2 1173 NA Li et al. (2002, 2004)
LaFeO3 6 11 92 4 1173 61 95 0 O2 11 92 1173 NA Dai et al. (2008a)

P.T. Krenzke et al. / Solar Energy xxx (2017) xxxxxx


66 99 0
69 96 0
La0.8Sr0.2FeO3 11 71 65 0
79 99 0
84 96 0
La0.7Sr0.3FeO3 2 100 0.5 ml g1 pulse 40 pulses 1273 59 42 0 H2O 100 Constant Volume Pulses 1.6 ml g1 1273 90 Nalbandian et al. (2009)
7.8 ml g1 20
17 ml g1 12
La0.3Sr0.7FeO3 40 59 0 1.6 ml g1 84
8.4 ml g1 20
12.9 ml g1 11
17.4 ml g1 7
SrFeO3 39 40 0 1.6 ml g1 75
5.9 ml g1 38
11.2 ml g1 17
16.8 ml g1 9
LaFeO3 100 0.5 ml g1 pulse 40 pulses 1273 11 64 0 H2O 100 Constant Volume Pulses 7.8 ml g1 1273 30 Evdou et al. (2010)
15.7 ml g1 22
22.4 ml g1 17
La0.7Sr0.3FeO3 32 43 0 4.5 ml g1 47
16.8 ml g1 23
22.4 ml g1 19
La0.3Sr0.7FeO3 39 40 0 6.7 ml g1 45
17.9 ml g1 34
22,4 ml g1 27
SrFeO3 33 29 0 5.6 ml g1 49
16.8 ml g1 32
22.4 ml g1 25
P.T. Krenzke et al. / Solar Energy xxx (2017) xxxxxx 19

The reaction mechanisms suggest how the composition and

Evdou et al. (2010)

Mudu et al. (2016)


morphology of the oxygen carrier material and cycling conditions

Zhao et al. (2016)

Dai et al. (2008a)


will influence methane and oxidizer conversion and syngas selec-
tivity. Changes to composition that cause oxygen to bond more
weakly with the metal oxide (e.g. doping of ceria) improve
methane conversion but reduce oxidizer conversion and syngas
Ref.

selectivity. High specific surface area yields higher methane and


oxidizer conversions and lower syngas selectivities (Otsuka et al.,
NA

NA
(K)

1998; Li et al., 2011b; Zheng et al., 2014b; Gao et al., 2016;


(min)

1123

1173

1273
Pantu et al., 2000; Evdou et al., 2010; Dai et al., 2008c). The rates

873
t

of the surface reactions and oxygen diffusion increase with increas-

100 pulses
ing temperature and thus higher temperature promotes higher
methane and oxidizer conversions (Krenzke et al., 2016; Li et al.,

ming
ml

2004, 2008b; Zhu et al., 2013; Pantu et al., 2000; Wei et al.,
20



V_

2008; Dai et al., 2006a).


Constant volume pulses
0.4 ml min1 g1 liquid water in 25

3.2.1. Ceria-based oxides


2.5 ml g1 pulses

Ceria has been studied widely as a redox material for thermo-


ml min1 g1 N2

chemical cycles in which reduction is carried out in an inert sweep


gas or at sub-atmospheric pressures (Lapp et al., 2012; Bader et al.,
2013, 2015; Krenzke and Davidson, 2015; Furler et al., 2012a,
92

2012b, 2014; Ermanoski et al., 2014; Bulfin et al., 2015;


(mol%)

Brendelberger et al., 2015; Venstrom et al., 2012, 2014; Lin and


100
Oxidation
yox

11

10

Haussener, 2015; Abanades and Flamant, 2006b; Kaneko et al.,


2007, 2008; Chueh and Haile, 2009, 2010; Chueh et al., 2010;
H2O
OX

O2

O2

Petkovich et al., 2011; Abanades et al., 2010; Meng et al., 2012;


(%)

Le Gal and Abanades, 2012; Scaranto and Idriss, 2015; Ramos-


24
12

18
39

15
21
Sc

8
0
0
0
0
0
0
0
0
0
0

Fernandez et al., 2014; Le Gal et al., 2011, 2013; Kang et al.,


SCO
(%)

22
48
43
34
14
65
99
96
27
96
38

43
41
51
51
49

2013, 2014; Dasari et al., 2013; Al-Shankiti et al., 2013; Scheffe



and Steinfeld, 2012; Rudisill et al., 2013; Lee et al., 2012; Jiang
(%)
SH2

62
53
52
53

45
33
43
44
61
66
69
50





et al., 2014b; Gibbons et al., 2014; Ackermann et al., 2014, 2015;


XCH4

Hathaway et al., 2015, 2016; Malonzo et al., 2014; Marxer et al.,


(%)

96
85
83
85
76
71
79
84
85
98
97
59
68
75
77
95

85
90

2015; Scheffe et al., 2014, 2013a; Takacs et al., 2015; Ji et al.,


1123

1173

1273

873

2016) and for CLR (Otsuka et al., 1993, 1998, 1999; Krenzke
(K)
T

et al., 2016; Warren et al., 2017; Sang et al., 2014; Sadykov et al.,
60 pulses

50 pulses

2004; Wei et al., 2007, 2010; Li et al., 2008a, 2008b, 2010a,


(min)

2010b, 2011a, 2011b, 2013b; Zhu et al., 2010, 2011a, 2011b,


20

2014, 2013; Gu et al., 2014, 2016; Kim et al., 2010; Jha et al.,
4
t

2015; Nair and Abanades, 2016; Gonzlez-Velasco et al., 1999;


2.5 ml g1 pulse
0.5 ml g1 pulse

Jiang et al., 2014a; Abanades and Flamant, 2006a, 2006b; He


et al., 2009; Gu et al., 2013; Jeong et al., 2011; Zheng et al.,
2014a, 2014b; Fathi et al., 2000; Gao et al., 2016; Pantu et al.,

ming
ml

2000). The fluorite structure is stable up to a nonstoichiometry of


25

92

V_
Reduction

0.25 (Chueh and Haile, 2010; Zinkevich et al., 2006; Kmmerle


(mol%)

and Heger, 1999), and the kinetics of oxygen diffusion are extre-
yCH4

100
11
40

10

mely fast (Chueh and Haile, 2010; Mogensen, 2000).


Krenzke et al. (2016) explored the impact of temperature and
5.187
2.254
2.419
2.503
3.480
10.57
 2

8.61
SSA

methane flow rate on methane conversion, syngas selectivity, and


m

oxidizer conversion with low surface area fibrous particles


0.20.5

(0.143 m2 g1). Methane and oxidizer conversion increased from


(mm)
Size

1173 K to 1273 K with modest improvements in syngas selectivity.







Carbon deposition was observed at 1373 K. They found a tradeoff


(Fe0.80Co0.20)0.75Al0.25O3
(Fe0.80Co0.20)0.9Al0.1O3

(Fe0.80Co0.20)0.6Al0.4O3
(Fe0.80Co0.20)0.4Al0.6O3

between the desire to maintain high methane conversion, high syn-


gas selectivity and high oxidizer conversion. Slower methane flow
(Fe0.80Co0.20) O3

rates increased the methane conversion but reduced syngas selec-


tivity and oxidizer conversion of oxidizer. Higher bed-average non-
stoichiometries were obtained with higher methane flow rates.
Warren et al. conducted limited experiments in a fixed bed reac-
10 wt.% NiO/La0.3Sr0.7FeO3
5 wt.% NiO/La0.3Sr0.7FeO3

tor in a solar simulator (Warren et al., 2017). The fixed bed held
Rh/La0.75Sr0.25
Rh/La0.75Sr0.25
Rh/La0.75Sr0.25
Rh/La0.75Sr0.25
Rh/La0.75Sr0.25
La0.8Sr0.2Fe0.9Co0.1O3

1130 g of ceria, which is 1/14th of the full capacity of the reactor.


SCLR with carbon dioxide was measured at 1393 K. Reaction (R4)
Table 8 (continued)

LaFe0.9Co0.1O3
LaFe0.7Co0.3O3
LaFe0.5Co0.5O3
LaFe0.3Co0.7O3

La0.8Sr0.2FeO3

was performed with methane diluted in argon (CH4:Ar = 1:9).


Composition

Methane conversion was 52%. Hydrogen and CO selectivities were


Material

0.5 wt.%
0.5 wt.%
0.5 wt.%
0.5 wt.%
0.5 wt.%
LaCoO3

83% and 59%, respectively. Carbon deposition was significant


(SC = 26%). Carbon dioxide conversion was 21%. The measured effi-
ciency, without accounting for separation work, was 0.92%. Warren

Please cite this article in press as: Krenzke, P.T., et al. Solar fuels via chemical-looping reforming. Sol. Energy (2017), http://dx.doi.org/10.1016/j.
solener.2017.05.095
20
solener.2017.05.095
Please cite this article in press as: Krenzke, P.T., et al. Solar fuels via chemical-looping reforming. Sol. Energy (2017), http://dx.doi.org/10.1016/j.

Table 9
AMnO3 perovskites.

Material Reduction Oxidation Ref.


Composition Size SSA yCH4 V_ t T XCH4 SH2 SCO Sc OX yox V_ t T Xox
 2    
(mm) m (mol%) ml (min) (K) (%) (%) (%) (%) (mol%) ml (min) (K) (%)
g ming ming

1 1
LaMnO3 2 100 0.5 ml g pulses 40 pulses 1273 20 24 42 0 H2O 100 Constant volume pulses 2.8 ml g 1273 38 Nalbandian et al. (2009) and
3.9 ml g1

P.T. Krenzke et al. / Solar Energy xxx (2017) xxxxxx


20 Evdou et al. (2008)
5.6 ml g1 12
La0.7Sr0.3MnO3 2 44 15 18 0 2.8 ml g1 32
3.9 ml g1 21
5.6 ml g1 16
La0.3Sr0.7MnO3 2 67 12 13 0 2.8 ml g1 40
3.9 ml g1 27
5.6 ml g1 19
CaMnO3 1.2 10 1000 20 1173 2 34 18 O2 10 1000 10 1173 NA Mishra et al. (2016)
BaMnO3 0.8 4 75 17
CaMn0.8Ni0.2O3 10 1000 20 1173 15 16 74 O2 10 1000 10 1173 NA Lin and Haussener (2015)
CaMn0.8Fe0.2O3 4 54 21
BaMnO3 0.8 10 1000 20 1173 4 75 17 O2 10 1000 10 1173 NA Mishra et al. (2016)
BaMn0.75Ni0.25O3 6 60 37
BaMn0.5Ni0.5O3 0.9 6 63 32
BaMnO3 0.8 10 1000 20 1173 4 75 17 O2 10 1000 10 1173 NA Mishra et al. (2016)
BaMn0.75Fe0.25O3 6 89 6
BaMn0.5Fe0.5O3 8 80 17
P.T. Krenzke et al. / Solar Energy xxx (2017) xxxxxx 21

et al. project efficiency would reach 7.5% if the reactions were per- 2014, 2013). Samples with separate and poorly dispersed Fe and Ce
formed at the full reactor capacity of 15.8 kg of ceria. Low efficiency phases with large crystal sizes produced by solid-phase synthesis
relative to thermodynamic predictions are due in part to operating exhibited lower methane conversion and higher syngas selectivity
with dilute methane, high thermal losses in a prototype, and prob- than samples prepared by hydrothermal and co-precipitation (Li
ably the selection of reactant flow rates and cycling times that did et al., 2010a). Compared to Zr4+ doped CeO2, Fe3+ doped CeO2 yields
not realize the optimum tradeoff in conversions and selectivities similar methane conversion but higher syngas selectivity and
for achieving high efficiency (Krenzke et al., 2016). lower carbon selectivity (Li et al., 2010b). Co-doping of Zr4+ and
Morphologies with higher specific surface area can improve Fe3+ provides a greater improvement in methane conversion while
syngas production rates if the surface area is stable over many maintaining or improving syngas selectivity (Li et al., 2011a).
redox cycles. Gao et al. (2016) found syngas production rates of
flame-made CeO2 (77 m2 g1) were as much as 167% higher than
3.2.2. Perovskites
that of commercial CeO2 fine powder (reported to be 8 m2 g1) at
Perovskites generally exhibit excellent redox properties, high
1173 K, but after 10 redox cycles 89% of the specific surface area
oxygen mobility, and thermal stability, and have been studied for
and 96% of the pore volume were lost, resulting in a 40% drop in
thermochemical water splitting (Balachandran et al., 1995;
syngas production. In agreement with prior studies of high surface
Choudhary et al., 2000; Trofimenko and Ullmann, 2000;
area three-dimensionally ordered macroporous (3DOM) and elec-
Vandenbossche and Mcintosh, 2008; Khanfekr et al., 2009; Haile,
trospun fiber ceria-based materials for swept gas reduction
2003; OConnell et al., 1999; Scheffe et al., 2013b; McDaniel
(Petkovich et al., 2011; Gibbons et al., 2014; Venstrom et al.,
et al., 2014, 2013; Demont et al., 2014; He and Li, 2015) and CLR
2012; Malonzo et al., 2014), sintering and loss of surface is a con-
(Rydn et al., 2008; Nalbandian et al., 2009, 2011; Evdou et al.,
cern even at 1200 K.
2008, 2010; Li et al., 2002, 2004; He et al., 2013a, 2013b; Dai
Promoting CeO2 with a Pt and Rh improves methane conversion
et al., 2008a, 2008b, 2008c, 2016, 2006a, 2006b; Wei et al., 2008;
without detriment to syngas selectivity or oxidizer conversion
Zhao et al., 2014, 2016; Mihai et al., 2012; Mudu et al., 2016;
(Otsuka et al., 1993, 1997, 1998; Fathi et al., 2000; Pantu et al.,
Mishra et al., 2016; Zeng et al., 2003; Garca et al., 2010;
2000). Catalysis is anticipated to occur via a reverse spillover
Shafiefarhood et al., 2015; Franca et al., 2012; Jiang et al., 2008;
mechanism in which H atoms trapped on Ce3+ spill over to a metal
Michalsky et al., 2015). La1xSrxB1yMyO3 with Fe and Mn B-site
site where the recombination and/or desorption occur more read-
cations have attracted the most attention for CLR. La1xSrxFeO3
ily (Otsuka et al., 1998). Methane conversion increased from 14% to
yields higher syngas selectivity and oxidizer than La1xSrxMnO3,
95% for both Pt and Rh. Carbon deposition and CO selectivity were
but methane conversion is lower (Nalbandian et al., 2009; Evdou
lower with Pt (Fathi et al., 2000).
et al., 2008, 2010). However neither material is fully reoxidized
Another approach for improving the methane activity and stabil-
with water because water splitting is thermodynamically unfavor-
ity of CeO2 is the partial substitution of Ce4+ ions with Zr4+ (Zhu et al.,
able for Fe and Mn B-site cations from the trivalent and to the
2011a; Gonzlez-Velasco et al., 1999; Otsuka et al., 1999; Jeong et al.,
tetravalent state (B3+ ? B4+) (Nalbandian et al., 2009, 2011).
2011; Li et al., 2010b; Zheng et al., 2014b, 2014a; Pantu et al., 2000;
Partial substitution of the A- and B-site cations can be used to
Jiang et al., 2014b). ZrO2 has a fluorite structure that permits Ce1x-
tailor the redox behavior (Rydn et al., 2008; Nalbandian et al.,
ZrxO2 solid solutions with Zr4+ content up to 40% (x = 0.4) (Kim,
2009; Evdou et al., 2008, 2010; Li et al., 2002, 2004; He et al.,
1989). The oxygen vacancy formation enthalpy and entropy
2013a; Dai et al., 2008a; Wei et al., 2008). Partial substitution of
decrease with increasing Zr content up to 20% (Scheffe et al.,
A-site cations in La1xSrxBO3 (B = Fe, Mn) increases the availability
2013a; Hao et al., 2014; Zhou et al., 2007). However, in accordance
of highly reactive oxygen (He et al., 2013a; Wei et al., 2008; Li et al.,
with the Evans-Polanyi principle, a reduction in the oxygen vacancy
2004). With increasing Sr content, methane conversion increases
formation energies also reduces the thermodynamic driving force
and oxidizer conversion and syngas selectivity decrease (Rydn
for oxidation, creating a tradeoff between higher methane conver-
et al., 2008; Nalbandian et al., 2009; Evdou et al., 2008, 2010; Li
sion with doped CeO2 and higher syngas selectivity and oxidizer
et al., 2002, 2004; He et al., 2013a; Dai et al., 2008a; Wei et al.,
conversion with undoped CeO2. Otsuka et al. (1999) measured the
2008). Partial substitution of B-site Fe in La1xSrxFe1yMyO3 has
impact of Zr4+ doping on the reactivity of Ce1xZrxO2 (x = 00.8)
been considered with M = Co, Ni, Cu, and Cr. Cr appears to be the
for methane partial oxidation at 973 K. Syngas production increased
most promising B-site dopant (Nalbandian et al., 2011). Methane
twofold for Zr4+ content up to 20% but was lower than that of pure
and oxidizer conversion increase with no change in hydrogen
CeO2 with higher Zr4+ content. The activation energy decreased from
selectivity for Cr content < 50%. The improvement in redox perfor-
137 kJ mol1 for CeO2 to 71 kJ mol1 for Ce0.8Zr0.2O2. Zhu et al.
mance with Cr doping was attributed to a change in oxidation state
(2011a) also observed an increase in methane conversion from 51
of Cr (Cr3+ to Cr6+) upon partial substitution of the A-site La3+ with
to 60% for partial Zr4+ substitution up to 30%. However, the increase
Sr2+ (Nalbandian et al., 2011). For AMn1yMyO3 (A = Ca, Ba), Ni3+
in conversion was accompanied by a decrease in H2 selectivity and in
and Fe3+ B-site dopants increased methane conversion, and Fe also
water splitting performance. This finding is consistent with non-CLR
increased syngas selectivity for Fe content up to 25% (y = 0.25)
studies in which oxidizer conversion was lower with Ce1xZrxO2
(Mishra et al., 2016).
than pure CeO2 (Gibbons et al., 2014; Bulfin et al., 2016). For Pt pro-
There are limited data for catalyst promoted perovskites (Evdou
moted Ce1xZrxO2, Zr4+, doping had no impact on methane conver-
et al., 2010; Nalbandian et al., 2011; Mudu et al., 2016). Addition of
sion but decreased H2 and CO selectivities for Zr4+ content up to
5 and 10 wt.% NiO to La1xSrxFeO3 provided an improvement in
50% (Pantu et al., 2000).
methane conversion and syngas selectivity and did not have an
Doping CeO2 with trivalent cations (e.g. Fe3+) can also influence
adverse effect on water splitting (Evdou et al., 2010; Nalbandian
redox properties (Scheffe and Steinfeld, 2012; Jiang et al., 2014b;
et al., 2011). There is minimal effect from increasing the NiO load-
Yahiro et al., 1988; Meng et al., 2011). The general finding is
ing from 5 wt.% to 10 wt.%.
methane conversion increases with Fe3+ content up to some critical
value (Li et al., 2008a, 2010a; He et al., 2009; Gu et al., 2013; Zhu
et al., 2013), and syngas selectivity decreases with increasing 4. Conclusion
Fe3+ (Li et al., 2008a, 2010a; He et al., 2009). Methane conversion
and syngas selectivity were stable for Ce1xFexO2 (x = 0.3, 0.5) over A review of the prior work on solar chemical looping reforming
525 redox cycles (Li et al., 2008b, 2010a, 2010b, 2011b; Zhu et al., (SCLR) to produce synthesis gas and split water or carbon dioxide

Please cite this article in press as: Krenzke, P.T., et al. Solar fuels via chemical-looping reforming. Sol. Energy (2017), http://dx.doi.org/10.1016/j.
solener.2017.05.095
22 P.T. Krenzke et al. / Solar Energy xxx (2017) xxxxxx

has been presented. In SCLR, synthesis gas and hydrogen (or carbon Abanades, S., Legal, A., Cordier, A., Peraudeau, G., Flamant, G., Julbe, A., 2010.
Investigation of reactive cerium-based oxides for H2 production by
monoxide) are produced in a two-step cycle. The introduction of
thermochemical two-step water-splitting. J. Mater. Sci. 45 (15), 41634173.
methane for the reduction of the metal oxide lowers the thermody- Ackermann, S., Scheffe, J.R., Steinfeld, A., 2014. Diffusion of oxygen in ceria at
namic barrier to split water or CO2 compared to reduction in an elevated temperatures and its application to H2O/CO2 splitting thermochemical
inert gas or vacuum. Syngas with a molar ratio of H2:CO equal to redox cycles. J. Phys. Chem. C 118 (10), 52165225.
Ackermann, S., Sauvin, L., Castiglioni, R., Rupp, J.L.M., Scheffe, J.R., Steinfeld, A., 2015.
2:1 comprises more than 99% of the equilibrium products at Kinetics of CO2 reduction over nonstoichiometric ceria. J. Phys. Chem. C 119
1223 K. The syngas can be used for downstream production of (29), 1645216461.
methanol or Fischer-Tropsch liquid fuels. The second step is reox- Adanez, J., Abad, A., Garcia-Labiano, F., Gayan, P., de Diego, L.F., 2012. Progress in
chemical-looping combustion and reforming technologies. Prog. Energy
idation of the metal oxide with steam to produce hydrogen or with Combust. Sci. 38 (2), 215282.
carbon dioxide to produce carbon monoxide. Water splitting is Agrafiotis, C., Roeb, M., Sattler, C., 2015. A review on solar thermal syngas
viewed as the most promising commercial option because of the production via redox pair-based water/carbon dioxide splitting
thermochemical cycles. Renew. Sustain. Energy Rev. 42, 254285.
flexibility of hydrogen for a variety of downstream processes and Al-Shankiti, I., Al-Otaibi, F., Al-Salik, Y., Idriss, H., 2013. Solar thermal hydrogen
because the energetic cost of producing a CO2 feedstock is high. production from water over modified CeO2 materials. Top. Catal. 56 (12), 1129
Moreover, hydrogen is easily separated from steam in the product 1138.
Bader, R., Venstrom, L.J., Davidson, J.H., Lipinski, W., 2013. Thermodynamic analysis
stream. With complete conversion of methane to syngas and water of isothermal redox cycling of ceria for solar fuel production. Energy Fuels 27
to hydrogen, the solar-to-chemical efficiency could be as high as (9), 55335544.
54% (using Eq. (2) and the parameters in Table 1). In this case, Bader, R., Bala Chandran, R., Venstrom, L.J., Sedler, S.J., Krenzke, P.T., De Smith, R.M.,
Banerjee, A., Chase, T.R., Davidson, J.H., Lipinski, W., 2015. Design of a solar
the energetic upgrade factor (Eq. (1)) would be 128%. Combustion
reactor to split CO2 via isothermal redox cycling of ceria. J. Sol. Energy Eng. 137
of the products would emit 41% less carbon per kilojoule than con- (3), 31007.
ventional non-solar CLR. Balachandran, U., Dusek, J.T., Mieville, R.L., Poeppel, R.B., Kleefisch, M.S., Pei, S.,
Key performance parameters to achieve the idealized efficiency Kobylinski, T.P., Udovich, C.A., Bose, A.C., 1995. Dense ceramic membranes for
partial oxidation of methane to syngas. Appl. Catal. A Gen. 133 (1), 1929.
are methane conversion, hydrogen selectivity, and carbon monox- Brendelberger, S., Roeb, M., Lange, M., Sattler, C., 2015. Counter flow sweep gas
ide selectivity for the partial oxidation reaction and oxidizer con- demand for the ceria redox cycle. Sol. Energy 122, 10111022.
version for water splitting. Efficiency is most sensitive to the Bulfin, B., Call, F., Lange, M., Lbben, O., Sattler, C., Pitz-Paal, R., Shvets, I.V., 2015.
Thermodynamics of CeO2 thermochemical fuel production. Energy Fuels 29 (2),
methane and oxidizer conversions. The projected efficiency of 10011009.
54% is an ideal case of conversions and selectivities of unity. For Bulfin, B., Call, F., Vieten, J., Roeb, M., Sattler, C., Shvets, I.V., 2016. Oxidation and
isothermal cycling, the penalty of lower conversions and selectivi- reduction reaction kinetics of mixed cerium zirconium oxides. J. Phys. Chem. C
120 (4), 20272035.
ties can be significant as shown in Figs. 10 and 11. Experimental Cha, K.-S., Kim, H.-S., Yoo, B.-K., Lee, Y.-S., Kang, K.-S., Park, C.-S., Kim, Y.-H., 2009.
studies of numerous stoichiometric and nonstoichiometric oxygen Reaction characteristics of two-step methane reforming over a Cu-ferrite/Ce
carrier material demonstrate there is often a tradeoff between ZrO2 medium. Int. J. Hydrogen Energy 34 (4), 18011808.
Cha, K.-S., Yoo, B.-K., Kim, H.-S., Ryu, T.-G., Kang, K.-S., Park, C.-S., Kim, Y.-H., 2010. A
attaining high methane conversion and high selectivities. Among study on improving reactivity of Cu-ferrite/ZrO2 medium for syngas and
the materials considered for CLR, ceria and perovskites that sup- hydrogen production from two-step thermochemical methane reforming. Int. J.
port high concentrations of oxygen vacancies, offer rapid diffusion Energy Res. 34 (5), 422430.
Cho, P., Mattisson, T., Lyngfelt, A., 2005. Carbon formation on nickel and iron oxide-
of oxygen and retain a stable crystal structure during cycling are
containing oxygen carriers for chemical-looping combustion. Ind. Eng. Chem.
promising. For partial oxidation of methane, platinum and rho- Res. 44 (4), 668676.
dium catalysts are beneficial. Of the perovskites, La1xSrxB1yMyO3 Choudhary, V., Banerjee, S., Uphade, B., 2000. Activation by hydrothermal treatment
with Fe and Mn B-site cations, but particularly La1xSrxFeO3 have of low surface area ABO3-type perovskite oxide catalysts. Appl. Catal. A Gen. 197
(2), L183L186.
been studied extensively. Chueh, W.C., Haile, S.M., 2009. Ceria as a thermochemical reaction medium for
There are limitations in the scope of the material studies that selectively generating syngas or methane from H2O and CO2. Chemsuschem 2
suggest avenues for future study. Continued study of perovskites (8), 735739.
Chueh, W.C., Haile, S.M., 2010. A thermochemical study of ceria: exploiting an old
can provide a more complete characterization of reoxidation for material for new modes of energy conversion and CO2 mitigation. Philos. Trans.
water splitting. In addition, data over broader and more practical R. Soc. A Math. Phys. Eng. Sci. 368 (1923), 32693294.
ranges of operating conditions will help identify candidate materi- Chueh, W.C., Falter, C., Abbott, M., Scipio, D., Furler, P., Haile, S.M., Steinfeld, A.,
2010, High-flux solar-driven thermochemical dissociation of CO2 and H2O using
als. Information is also lacking on the impact of material composi- nonstoichiometric ceria, Science (80-.)., 330(6012), pp. 17971801.
tion on carbon deposition and the stability of materials over many Dai, X.P., Li, R.J., Yu, C.C., Hao, Z.P., 2006a. Unsteady-state direct partial oxidation of
cycles. methane to synthesis gas in a fixed-bed reactor using AFeO3 (A = La, Nd, Eu)
perovskite-type oxides as oxygen storage. J. Phys. Chem. B 110 (45), 22525
The potential for high efficiency and flexibility of the products 22531.
for solar CLR supports development of materials and reactors Dai, X.P., Wu, Q., Li, R.J., Yu, C.C., Hao, Z.P., 2006b. Hydrogen production from a
designed specifically for SCLR as a step toward a more complete combination of the watergas shift and redox cycle process of methane partial
oxidation via lattice oxygen over LaFeO3 perovskite catalyst. J. Phys. Chem. B
assessment of feasibility in comparison to other routes for solar
110 (51), 2585625862.
fuels, particularly routes that utilize methane as feedstock. Dai, X., Yu, C., Wu, Q., 2008a. Comparison of LaFeO3, La0.8Sr0.2FeO3, and
La0.8Sr0.2Fe0.9Co0.1O3 perovskite oxides as oxygen carrier for partial oxidation
of methane. J. Nat. Gas Chem. 17 (4), 415418.
Acknowledgement Dai, X., Yu, C., Li, R., Wu, Q., Hao, Z., 2008b. Synthesis gas production using oxygen
storage materials as oxygen carrier over circulating fluidized bed. J. Rare Earths
26 (1), 7680.
Jesse R. Fosheim is supported by the National Science Founda- Dai, X., Yu, C., Li, R., Wu, Q., Shi, K., Hao, Z., 2008c. Effect of calcination temperature
tion Graduate Research Fellowship Program (NSF-GRFP) under and reaction conditions on methane partial oxidation using lanthanum-based
perovskite as oxygen donor. J. Rare Earths 26 (3), 341346.
Grant 00039202. Dai, X., Cheng, J., Li, Z., Liu, M., Ma, Y., Zhang, X., 2016. Reduction kinetics of
lanthanum ferrite perovskite for the production of synthesis gas by chemical-
looping methane reforming. Chem. Eng. Sci. 153, 236245.
References Dasari, H.P., Ahn, K., Park, S.-Y., Ji, H.-I., Yoon, K.J., Kim, B.-K., Je, H.-J., Lee, H.-W., Lee,
J.-H., 2013. Hydrogen production from water-splitting reaction based on RE-
doped ceriazirconia solid-solutions. Int. J. Hydrogen Energy 38 (14), 6097
Abanades, S., Flamant, G., 2006a. Solar hydrogen production from the thermal
6103.
splitting of methane in a high temperature solar chemical reactor. Sol. Energy
de Diego, L.F., Ortiz, M., Adnez, J., Garca-Labiano, F., Abad, A., Gayn, P., 2008.
80 (10), 13211332.
Synthesis gas generation by chemical-looping reforming in a batch fluidized
Abanades, S., Flamant, G., 2006b. Thermochemical hydrogen production from a
bed reactor using Ni-based oxygen carriers. Chem. Eng. J. 144 (2), 289298.
two-step solar-driven water-splitting cycle based on cerium oxides. Sol. Energy
de Klerk, A., 2011. Fischer-Tropsch Refining. John Wiley & Sons Inc.
80 (12), 16111623.

Please cite this article in press as: Krenzke, P.T., et al. Solar fuels via chemical-looping reforming. Sol. Energy (2017), http://dx.doi.org/10.1016/j.
solener.2017.05.095
P.T. Krenzke et al. / Solar Energy xxx (2017) xxxxxx 23

Demont, A., Abanades, S., Beche, E., 2014. Investigation of perovskite structures as Hao, Y., Yang, C.-K., Haile, S.M., 2014. Ceriazirconia solid solutions (Ce1XZrXO2d,
oxygen-exchange redox materials for hydrogen production from X 0.2) for solar thermochemical water splitting: a thermodynamic study.
thermochemical two-step water-splitting cycles. J. Phys. Chem. C 118 (24), Chem. Mater. 26 (20), 60736082.
1268212692. Hathaway, B.J., Bala Chandran, R., Sedler, S., Thomas, D., Gladen, A., Chase, T.,
Dueso, C., Ortiz, M., Abad, A., Garca-Labiano, F., de Diego, L.F., Gayn, P., Adnez, J., Davidson, J.H., 2015. Effect of flow rates on operation of a solar thermochemical
2012. Reduction and oxidation kinetics of nickel-based oxygen-carriers for reactor for splitting CO2 via the isothermal ceria redox cycle. J. Sol. Energy Eng.
chemical-looping combustion and chemical-looping reforming. Chem. Eng. J. 138 (1), 11007.
188, 142154. Hathaway, B.J., Bala Chandran, R., Gladen, A.C., Chase, T.R., Davidson, J.H., 2016.
Ehrhart, B.D., Muhich, C.L., Al-Shankiti, I., Weimer, A.W., 2016a. System efficiency Demonstration of a solar reactor for carbon dioxide splitting via the isothermal
for two-step metal oxide solar thermochemical hydrogen production part 2: ceria redox cycle and practical implications. Energy Fuels 30 (8), 66546661.
impact of gas heat recuperation and separation temperatures. Int. J. Hydrogen He, F., Li, F., 2015. Perovskite promoted iron oxide for hybrid water-splitting and
Energy, 110. syngas generation with exceptional conversion. Energy Environ. Sci. 8 (2), 535
Ehrhart, B.D., Muhich, C.L., Al-Shankiti, I., Weimer, A.W., 2016b. System efficiency 539.
for two-step metal oxide solar thermochemical hydrogen production Part 1: He, F., Wei, Y., Li, H., Wang, H., 2009. Synthesis gas generation by chemical-looping
Thermodynamic model and impact of oxidation kinetics. Int. J. Hydrogen reforming using Ce-based oxygen carriers modified with Fe, Cu, and Mn oxides.
Energy 41 (44), 1988119893. Energy Fuels 23 (4), 20952102.
Ehrhart, B.D., Muhich, C.L., Al-Shankiti, I., Weimer, A.W., 2016c. System efficiency He, F., Li, X., Zhao, K., Huang, Z., Wei, G., Li, H., 2013a. The use of La1xSrxFeO3
for two-step metal oxide solar thermochemical hydrogen production Part 3: perovskite-type oxides as oxygen carriers in chemical-looping reforming of
Various methods for achieving low oxygen partial pressures in the reduction methane. Fuel 108, 465473.
reaction. Int. J. Hydrogen Energy 41 (44), 1990419914. He, F., Zhao, K., Huang, Z., Li, X., Wei, G., Li, H., 2013b. Synthesis of three-
U.S. Energy Information Administration (EIA), March 2016 Month Energy Review. dimensionally ordered macroporous LaFeO3 perovskites and their performance
U.S. Energy Information Administration (EIA), International Energy Outlook, 2016. for chemical-looping reforming of methane. Chinese J. Catal. 34, 12421249.
Ermanoski, I., 2014. Cascading pressure thermal reduction for efficient solar fuel Herron, J.A., Kim, J., Upadhye, A.A., Huber, G.W., Maravelias, C.T., 2015. A general
production. Int. J. Hydrogen Energy 39 (25), 1311413117. framework for the assessment of solar fuel technologies. Energy Environ. Sci. 8
Ermanoski, I., Siegel, N.P., Stechel, E.B., 2013. A new reactor concept for (1), 126157.
efficient solar-thermochemical fuel production. J. Sol. Energy Eng. 135 (3), House, K.Z., Baclig, A.C., Ranjan, M., van Nierop, E.a., Wilcox, J., Herzog, H.J., 2011.
31002. Economic and energetic analysis of capturing CO2 from ambient air. Proc. Natl.
Ermanoski, I., Miller, J.E., Allendorf, M.D., 2014. Efficiency maximization in solar- Acad. Sci. 108 (51), 2042820433.
thermochemical fuel production: challenging the concept of isothermal water Jang, J.T., Kwak, J.H., Han, G.Y., Bae, J.W., Ahn, C.-I., Yoon, K.J., 2015. Redox of
splitting. Phys. Chem. Chem. Phys. 16 (18), 8418. titanium oxides by methane and water for application to cyclic syngas and
Evdou, A., Zaspalis, V., Nalbandian, L., 2008. La(1x)SrxMnO3d perovskites as redox hydrogen production systems. Int. J. Hydrogen Energy 40 (6), 25182528.
materials for the production of high purity hydrogen. Int. J. Hydrogen Energy 33 Jeong, H.H., Kwak, J.H., Han, G.Y., Yoon, K.J., 2011. Stepwise production of syngas
(20), 55545562. and hydrogen through methane reforming and water splitting by using a
Evdou, A., Zaspalis, V., Nalbandian, L., 2010. La1xSrxFeO3d perovskites as redox cerium oxide redox system. Int. J. Hydrogen Energy 36 (23), 1522115230.
materials for application in a membrane reactor for simultaneous production of Jha, A., Jeong, D.-W., Jang, W.-J., Lee, Y.-L., Roh, H.-S., 2015. Hydrogen production
pure hydrogen and synthesis gas. Fuel 89 (6), 12651273. from watergas shift reaction over NiCuCeO2 oxide catalyst: the effect of
Fathi, M., Bjorgum, E., Viig, T., Rokstad, O., 2000. Partial oxidation of methane to preparation methods. Int. J. Hydrogen Energy 40 (30), 92099216.
synthesis gas. Catal. Today 63 (24), 489497. Ji, H.-I., Davenport, T.C., Gopal, C.B., Haile, S.M., 2016. Extreme high temperature
Forster, M., 2004. Theoretical investigation of the system SnOx/Sn for the redox kinetics in ceria: exploration of the transition from gas-phase to material-
thermochemical storage of solar energy. Energy 29 (56), 789799. kinetic limitations. Phys. Chem. Chem. Phys. 18 (31), 2155421561.
Franca, R.V., Thursfield, A., Metcalfe, I.S., 2012. La0.6Sr0.4Co0.2Fe0.8O3d microtubular Jiang, H., Wang, H., Werth, S., Schiestel, T., Caro, J., 2008. Simultaneous production of
membranes for hydrogen production from water splitting. J. Memb. Sci. 389, hydrogen and synthesis gas by combining water splitting with partial oxidation
173181. of methane in a hollow-fiber membrane reactor. Angew. Chem. Int. Ed. 47 (48),
Furler, P., Scheffe, J.R., Steinfeld, A., 2012a. Syngas production by simultaneous 93419344.
splitting of H2O and CO2 via ceria redox reactions in a high-temperature solar Jiang, Q., Zhou, G., Jiang, Z., Li, C., 2014a. Thermochemical CO2 splitting reaction with
reactor. Energy Environ. Sci. 5 (3), 60986103. CexM1xO2d (M = Ti4+, Sn4+, Hf4+, Zr4+, La3+, Y3+ and Sm3+) solid solutions. Sol.
Furler, P., Scheffe, J., Gorbar, M., Moes, L., Vogt, U., Steinfeld, A., 2012b. Solar Energy 99, 5566.
thermochemical CO2 splitting utilizing a reticulated porous ceria redox system. Jiang, Q., Zhou, G., Jiang, Z., Li, C., 2014b. Thermochemical CO2 splitting reaction
Energy Fuels 26 (11), 70517059. with CexM1xO2d (M = Ti4+, Sn4+, Hf4+, Zr4+, La3+, Y3+ and Sm3+) solid solutions.
Furler, P., Scheffe, J., Marxer, D., Gorbar, M., Bonk, A., Vogt, U., Steinfeld, A., 2014. Sol. Energy 99, 5566.
Thermochemical CO2 splitting via redox cycling of ceria reticulated foam Kaneko, H., Miura, T., Ishihara, H., Taku, S., Yokoyama, T., Nakajima, H., Tamaura, Y.,
structures with dual-scale porosities. Phys. Chem. Chem. Phys. 16 (22), 10503 2007. Reactive ceramics of CeO2MOx (M = Mn, Fe, Ni, Cu) for H2 generation by
10511. two-step water splitting using concentrated solar thermal energy. Energy 32
Gao, X., Vidal, A., Bayon, A., Bader, R., Hinkley, J., Lipinski, W., Tricoli, A., 2016. (5), 656663.
Efficient ceria nanostructures for enhanced solar fuel production via high- Kaneko, H., Ishihara, H., Taku, S., Naganuma, Y., Hasegawa, N., Tamaura, Y., 2008.
temperature thermochemical redox cycles. J. Mater. Chem. A 4 (24), 9614 Cerium ion redox system in CeO2xFe2O3 solid solution at high temperatures
9624. (12731673 K) in the two-step water-splitting reaction for solar H2 generation.
Garca, V., Caldes, M.T., Joubert, O., Gautron, E., Mondragn, F., Moreno, A., 2010. J. Mater. Sci. 43 (9), 31533161.
Methane oxidation by lattice oxygen of Ni/BaTi1xInxO3d catalysts. Catal. Today Kang, K.-S., Kim, C.-H., Cho, W.-C., Bae, K.-K., Woo, S.-W., Park, C.-S., 2008. Reduction
157 (14), 177182. characteristics of CuFe2O4 and Fe3O4 by methane; CuFe2O4 as an oxidant for
Gibbons, W.T., Venstrom, L.J., De Smith, R.M., Davidson, J.H., Jackson, G.S., 2014. two-step thermochemical methane reforming. Int. J. Hydrogen Energy 33 (17),
Ceria-based electrospun fibers for renewable fuel production via two-step 45604568.
thermal redox cycles for carbon dioxide splitting. Phys. Chem. Chem. Phys. 16 Kang, K.-S., Kim, C.-H., Bae, K.-K., Cho, W.-C., Kim, W.-J., Kim, Y.-H., Kim, S.-H., Park,
(27), 14271. C.-S., 2010. Redox cycling of CuFe2O4 supported on ZrO2 and CeO2 for two-step
Go, K., Son, S., Kim, S., 2008. Reaction kinetics of reduction and oxidation methane reforming/water splitting. Int. J. Hydrogen Energy 35 (2), 568576.
of metal oxides for hydrogen production. Int. J. Hydrogen Energy 33 (21), Kang, M., Zhang, J., Wang, C., Wang, F., Zhao, N., Xiao, F., Wei, W., Sun, Y., 2013. CO2
59865995. splitting via two step thermochemical reactions over doped ceria/zirconia solid
Go, K., Son, S., Kim, S., Kang, K., Park, C., 2009. Hydrogen production from two-step solutions. RSC Adv. 3 (41), 18878.
steam methane reforming in a fluidized bed reactor. Int. J. Hydrogen Energy 34 Kang, M., Wu, X., Zhang, J., Zhao, N., Wei, W., Sun, Y., 2014. Enhanced
(3), 13011309. thermochemical CO2 splitting over Mg- and Ca-doped ceria/zirconia solid
Gonzlez-Velasco, J.R., Gutirrez-Ortiz, M.A., Marc, J.-L., Botas, J.A., Gonzlez- solutions. RSC Adv. 4 (11), 5583.
Marcos, M.P., Blanchard, G., 1999. Contribution of cerium/zirconium mixed Khanfekr, a., Arzani, K., Nemati, A., Hosseini, M., 2009. Production of perovskite
oxides to the activity of a new generation of TWC. Appl. Catal. B Environ. 22 (3), catalysts on ceramic monoliths with nanoparticles for dual fuel system
167178. automobiles. Int. J. Environ. Sci. Technol. 6 (1), 105112.
Gu, Z., Li, K., Wang, H., Wei, Y., Yan, D., Qiao, T., 2013. Syngas production from Kim, D., 1989. Lattice parameters, ionic conductivities, and solubility limits in
methane over CeO2-Fe2O3 mixed oxides using a chemical-looping method. fluorite-structure MO2 oxide [M = Hf4+, Zr4+, Ce4+, Th4+, U4+] solid solutions. J.
Kinet. Catal. 54 (3), 326333. Am. Ceram. Soc. 72 (8), 14151421.
Gu, Z., Li, K., Qing, S., Zhu, X., Wei, Y., Li, Y., Wang, H., 2014. Enhanced reducibility Kim, H.-S., Cha, K.-S., Yoo, B.-K., Ryu, T.-G., Lee, Y.-S., Park, C.-S., Kim, Y.-H., 2010.
and redox stability of Fe2O3 in the presence of CeO2 nanoparticles. RSC Adv. 4 Chemical hydrogen storage and release properties using redox reaction over the
(88), 4719147199. Cu-added Fe/Ce/Zr mixed oxide medium. J. Ind. Eng. Chem. 16 (1), 8186.
Gu, Z., Li, K., Wang, H., Qing, S., Zhu, X., Wei, Y., Cheng, X., Yu, H., Cao, Y., 2016. Bulk Kim, J., Henao, C.A., Johnson, T.A., Dedrick, D.E., Miller, J.E., Stechel, E.B., Maravelias,
monolithic CeZrFeO/Al2O3 oxygen carriers for a fixed bed scheme of the C.T., 2011. Methanol production from CO2 using solar-thermal energy: process
chemical looping combustion: reactivity of oxygen carrier. Appl. Energy 163, development and techno-economic analysis. Energy Environ. Sci. 4 (9), 3122.
1931. Kim, J., Johnson, T.A., Miller, J.E., Stechel, E.B., Maravelias, C.T., 2012. Fuel production
Haile, S.M., 2003. Fuel cell materials and components. Acta Mater. 51 (19), 5981 from CO2 using solar-thermal energy: system level analysis. Energy Environ. Sci.
6000. 5 (9), 8417.

Please cite this article in press as: Krenzke, P.T., et al. Solar fuels via chemical-looping reforming. Sol. Energy (2017), http://dx.doi.org/10.1016/j.
solener.2017.05.095
24 P.T. Krenzke et al. / Solar Energy xxx (2017) xxxxxx

Kodama, T., 2000. Thermochemical methane reforming using a reactive WO3/W Meng, Q.-L., Lee, C., Ishihara, T., Kaneko, H., Tamaura, Y., 2011. Reactivity of CeO2-
redox system. Energy 25 (5), 411425. based ceramics for solar hydrogen production via a two-step water-splitting
Kodama, T., Shimizu, T., Satoh, T., Nakata, M., Shimizu, K.-I., 2002. Stepwise cycle with concentrated solar energy. Int. J. Hydrogen Energy 36 (21), 13435
production of CO-rich syngas and hydrogen via solar methane reforming by 13441.
using a Ni(II)ferrite redox system. Sol. Energy 73 (5), 363374. Meng, Q.-L., Lee, C., Shigeta, S., Kaneko, H., Tamaura, Y., 2012. Solar hydrogen
Kodama, T., Shimizu, T., Satoh, T., Shimizu, K.-I., 2003. Stepwise production of CO- production using Ce1xLixO2d solid solutions via a thermochemical, two-step
rich syngas and hydrogen via methane reforming by a WO3-redox catalyst. water-splitting cycle. J. Solid State Chem. 194, 343351.
Energy 28 (11), 10551068. Michalsky, R., Neuhaus, D., Steinfeld, A., 2015. Carbon dioxide reforming of methane
Krupl, S., Steinfeld, A., 2001a. Pulsed gas feeding for stoichiometric operation of a using an isothermal redox membrane reactor. Energy Technol. 3 (7), 784789.
gas-solid vortex flow solar chemical reactor. J. Sol. Energy Eng. 123 (2), 133. Mihai, O., Chen, D., Holmen, A., 2012. Chemical looping methane partial oxidation:
Krupl, S., Steinfeld, A., 2001b. Experimental investigation of a vortex-flow solar the effect of the crystal size and O content of LaFeO3. J. Catal. 293, 175185.
chemical reactor for the combined ZnO-reduction and CH[sub 4]-reforming. J. Miller, J.E., McDaniel, A.H., Allendorf, M.D., 2014. Considerations in the design of
Sol. Energy Eng. 123 (3), 237. materials for solar-driven fuel production using metal-oxide thermochemical
Krenzke, P.T., Davidson, J.H., 2014. Thermodynamic analysis of syngas production cycles. Adv. Energy Mater. 4 (2), 1300469.
via the solar thermochemical cerium oxide redox cycle with methane-driven Mishra, A., Galinsky, N., He, F., Santiso, E.E., Li, F., 2016. Perovskite-structured
reduction. Energy Fuels 28 (6), 40884095. AMnxB1x O3 (A = Ca or Ba; B = Fe or Ni) redox catalysts for partial oxidation of
Krenzke, P.T., Davidson, J.H., 2015. On the efficiency of solar H2 and CO production methane. Catal. Sci. Technol. 6 (12), 45354544.
via the thermochemical cerium oxide redox cycle: the option of inert-swept Mogensen, M., 2000. Physical, chemical and electrochemical properties of pure and
reduction. Energy Fuels 29 (2), 10451054. doped ceria. Solid State Ionics 129 (14), 6394.
Krenzke, P.T., Fosheim, J.R., Zheng, J., Davidson, J.H., 2016. Synthesis gas production Mudu, F., Olsbye, U., Arstad, B., Diplas, S., Li, Y., Fjellvg, H., 2016. Aluminium
via the solar partial oxidation of methane-ceria redox cycle: conversion, substituted lanthanum based perovskite type oxides, non-stoichiometry and
selectivity, and efficiency. Int. J. Hydrogen Energy 41 (30), 1279912811. performance in methane partial oxidation by framework oxygen. Appl. Catal. A
Kmmerle, E., Heger, G., 1999. The structures of C-Ce2O3+d, Ce7O12, and Ce11O20. J. Gen. 523, 171181.
Solid State Chem. 147 (2), 485500. Muhich, C.L., Evanko, B.W., Weston, K.C., Lichty, P., Liang, X., Martinek, J., Musgrave,
Lapp, J., Davidson, J.H., Lipinski, W., 2012. Efficiency of two-step solar C.B., Weimer, A.W., 2013. Efficient generation of H2 by splitting water with an
thermochemical non-stoichiometric redox cycles with heat recovery. Energy isothermal redox cycle. Science (80-.)., 341(6145), pp. 540542.
37 (1), 591600. Muhich, C.L., Ehrhart, B.D., Al-Shankiti, I., Ward, B.J., Musgrave, C.B., Weimer, A.W.,
Le Gal, A., Abanades, S., 2012. Dopant incorporation in ceria for enhanced water- 2016. A review and perspective of efficient hydrogen generation via solar
splitting activity during solar thermochemical hydrogen generation. J. Phys. thermal water splitting. Wiley Interdiscip. Rev. Energy Environ. 5 (3), 261287.
Chem. C 116 (25), 1351613523. Nair, M.M., Abanades, S., 2016. Tailoring hybrid nonstoichiometric ceria redox cycle
Le Gal, A., Abanades, S., Flamant, G., 2011. CO2 and H2O splitting for thermochemical for combined solar methane reforming and thermochemical conversion of H2O/
production of solar fuels using nonstoichiometric ceria and ceria/zirconia solid CO2. Energy Fuels 30 (7), 60506058.
solutions. Energy Fuels 25 (10), 48364845. Nakayama, O., Ikenaga, N., Miyake, T., Yagasaki, E., Suzuki, T., 2010. Production of
Le Gal, A., Abanades, S., Bion, N., Le Mercier, T., Harl, V., 2013. Reactivity of doped synthesis gas from methane using lattice oxygen of NiOCr2O3MgO complex
ceria-based mixed oxides for solar thermochemical hydrogen generation via oxide. Ind. Eng. Chem. Res. 49 (2), 526534.
two-step water-splitting cycles. Energy Fuels 27 (10), 60686078. Nalbandian, L., Evdou, A., Zaspalis, V., 2009. La1xSrxMO3 (M = Mn, Fe) perovskites
Lee, C., Meng, Q.-L., Kaneko, H., Tamaura, Y., 2012. Solar hydrogen productivity of as materials for thermochemical hydrogen production in conventional and
ceria-scandia solid solution using two-step water-splitting cycle. J. Sol. Energy membrane reactors. Int. J. Hydrogen Energy 34 (17), 71627172.
Eng. 135 (1), 11002. Nalbandian, L., Evdou, A., Zaspalis, V., 2011. La1xSrxMyFe1yO3d perovskites as
Li, R., Yu, C., Shen, S., 2002. Partial oxidation of methane to syngas using lattice oxygen-carrier materials for chemical-looping reforming. Int. J. Hydrogen
oxygen of La1xSrxFeO3 perovskite oxide catalysts instead of molecular oxygen. Energy 36 (11), 66576670.
J. Nat. Gas Chem. 11, 137144. OConnell, M., Norman, A., Httermann, C., Morris, M., 1999. Catalytic oxidation
Li, R.-J., Yu, C.-C., Ji, W.-J., Shen, S.-K., 2004. Methane oxidation to synthesis gas over lanthanum-transition metal perovskite materials. Catal. Today 47 (14),
using lattice oxygen in La1xSrxFeO3 perovskite oxides instead of molecular 123132.
oxygen. Stud. Surf. Sci. Catal., 199204 Otsuka, K., Ushiyama, T., Yamanaka, I., 1993. Partial oxidation of methane using the
Li, K., Wang, H., Wei, Y., Liu, M., 2008a. Preparation and characterization of redox of cerium oxide. Chem. Lett. 22 (9), 15171520.
Ce1xFexO2 complex oxides and its catalytic activity for methane selective Otsuka, K., Sunada, E., Ushiyama, T., Yamanaka, I., 1997. The production of synthesis
oxidation. J. Rare Earths 26 (2), 245249. gas by the redox of cerium oxide. Stud. Surf. Sci. Catal. 107 (1), 531536.
Li, K., Wang, H., Wei, Y., Liu, M., 2008b. Catalytic performance of cerium iron Otsuka, K., Wang, Y., Sunada, E., Yamanaka, I., 1998. Direct partial oxidation of
complex oxides for partial oxidation of methane to synthesis gas. J. Rare Earths methane to synthesis gas by cerium oxide. J. Catal. 175 (2), 152160.
26 (5), 705710. Otsuka, K., Wang, Y., Nakamura, M., 1999. Direct conversion of methane to
Li, K., Wang, H., Wei, Y., Yan, D., 2010a. Direct conversion of methane to synthesis synthesis gas through gassolid reaction using CeO2ZrO2 solid solution at
gas using lattice oxygen of CeO2Fe2O3 complex oxides. Chem. Eng. J. 156 (3), moderate temperature. Appl. Catal. A Gen. 183 (2), 317324.
512518. Pantu, P., Kim, K., Gavalas, G.R., 2000. Methane partial oxidation on Pt/CeO2ZrO2 in
Li, K., Wang, H., Wei, Y., Yan, D., 2010b. Syngas production from methane and air via the absence of gaseous oxygen. Appl. Catal. A Gen. 193 (12), 203214.
a redox process using CeFe mixed oxides as oxygen carriers. Appl. Catal. B Petkovich, N.D., Rudisill, S.G., Venstrom, L.J., Boman, D.B., Davidson, J.H., Stein, A.,
Environ. 97 (34), 361372. 2011. Control of heterogeneity in nanostructured Ce1xZrxO2 binary oxides for
Li, K., Wang, H., Wei, Y., Yan, D., 2011a. Transformation of methane into synthesis enhanced thermal stability and water splitting activity. J. Phys. Chem. C 115
gas using the redox property of CeFe mixed oxides: effect of calcination (43), 2102221033.
temperature. Int. J. Hydrogen Energy 36 (5), 34713482. Protasova, L., Snijkers, F., 2016. Recent developments in oxygen carrier materials for
Li, K., Wang, H., Wei, Y., Yan, D., 2011b. Partial oxidation of methane to syngas with hydrogen production via chemical looping processes. Fuel 181, 7593.
air by lattice oxygen transfer over ZrO2-modified CeFe mixed oxides. Chem. Ramos-Fernandez, E.V., Shiju, N.R., Rothenberg, G., 2014. Understanding the solar-
Eng. J. 173 (2), 574582. driven reduction of CO2 on doped ceria. RSC Adv. 4 (32), 16456.
Li, K., Wang, H., Wei, Y., 2013a. Syngas generation from methane using a chemical- Roeb, M., Neises, M., Monnerie, N., Call, F., Simon, H., Sattler, C., Schmcker, M., Pitz-
looping concept: a review of oxygen carriers. J. Chem. 2013, 18. Paal, R., 2012. Materials-related aspects of thermochemical water and carbon
Li, K., Haneda, M., Gu, Z., Wang, H., Ozawa, M., 2013b. Modification of CeO2 on the dioxide splitting: a review. Materials (Basel) 5 (12), 20152054.
redox property of Fe2O3. Mater. Lett. 93, 129132. Rostrup-Nielsen, J.R., 2002. Syngas in perspective. Catal. Today 71 (34), 243247.
Lin, M., Haussener, S., 2015. Solar fuel processing efficiency for ceria redox cycling Rudisill, S.G., Venstrom, L.J., Petkovich, N.D., Quan, T., Hein, N., Boman, D.B.,
using alternative oxygen partial pressure reduction methods. Energy 88, 667 Davidson, J.H., Stein, A., 2013. Enhanced oxidation kinetics in thermochemical
679. cycling of CeO2 through templated porosity. J. Phys. Chem. C 117 (4), 1692
Malonzo, C.D., De Smith, R.M., Rudisill, S.G., Petkovich, N.D., Davidson, J.H., Stein, A., 1700.
2014. Wood-templated CeO2 as active material for thermochemical CO Rydn, M., Lyngfelt, A., Mattisson, T., Chen, D., Holmen, A., Bjrgum, E., 2008. Novel
production. J. Phys. Chem. C 118 (45), 2617226181. oxygen-carrier materials for chemical-looping combustion and chemical-
Mancini, T., Heller, P., Butler, B., Osborn, B., Schiel, W., Goldberg, V., Buck, R., Diver, looping reforming; LaxSr1xFeyCo1yO3d perovskites and mixed-metal oxides
R., Andraka, C., Moreno, J., 2003. Dish-stirling systems: an overview of of NiO, Fe2O3 and Mn3O4. Int. J. Greenh. Gas Control 2 (1), 2136.
development and status. J. Sol. Energy Eng. 125 (2), 135. Sadykov, V.A., Kuznetsova, T.G., Alikina, G.M., Frolova, Y.V., Lukashevich, A.I.,
Marxer, D., Furler, P., Scheffe, J., Geerlings, H., Falter, C., Batteiger, V., Sizmann, A., Potapova, Y.V., Muzykantov, V.S., Rogov, V.A., Kriventsov, V.V., Kochubei, D.I.,
Steinfeld, A., 2015. Demonstration of the entire production chain to renewable Moroz, E.M., Zyuzin, D.I., Zaikovskii, V.I., Kolomiichuk, V.N., Paukshtis, E.A.,
kerosene via solar thermochemical splitting of H2O and CO2. Energy Fuels 29 Burgina, E.B., Zyryanov, V.V., Uvarov, N.F., Neophytides, S., Kemnitz, E., 2004.
(5), 32413250. Ceria-based fluorite-like oxide solid solutions as catalysts of methane selective
McDaniel, A.H., Miller, E.C., Arifin, D., Ambrosini, A., Coker, E.N., OHayre, R., Chueh, oxidation into syngas by the lattice oxygen: synthesis, characterization and
W.C., Tong, J., 2013. Sr- and Mn-doped LaAlO3d for solar thermochemical H2 performance. Catal. Today 9395, 4553.
and CO production. Energy Environ. Sci. 6 (8), 2424. Said, S.A.M., Waseeuddin, M., Simakov, D.S.A., 2016. A review on solar reforming
McDaniel, A.H., Ambrosini, A., Coker, E.N., Miller, J.E., Chueh, W.C., OHayre, R., Tong, systems. Renew. Sustain. Energy Rev. 59, 149159.
J., 2014. Nonstoichiometric perovskite oxides for solar thermochemical H2 and Sang, X.-L., Li, K.-Z., Wang, H., Wei, Y.-G., 2014. Selective oxidation of methane and
CO production. Energy Proc. 49, 20092018. carbon deposition over Fe2O3/Ce1x Zr X O2 oxides. Rare Met. 33 (2), 230238.

Please cite this article in press as: Krenzke, P.T., et al. Solar fuels via chemical-looping reforming. Sol. Energy (2017), http://dx.doi.org/10.1016/j.
solener.2017.05.095
P.T. Krenzke et al. / Solar Energy xxx (2017) xxxxxx 25

Scaranto, J., Idriss, H., 2015. The effect of uranium cations on the redox properties of Takenaka, S., Nomura, K., Hanaizumi, N., Otsuka, K., 2005. Storage and formation of
CeO2 within the context of hydrogen production from water. Top. Catal. 58 (2 pure hydrogen mediated by the redox of modified iron oxides. Appl. Catal. A
3), 143148. Gen. 282 (12), 333341.
Scheffe, J.R., Steinfeld, A., 2012. Thermodynamic analysis of cerium-based oxides for Trofimenko, N.E., Ullmann, H., 2000. Oxygen stoichiometry and mixed ionic-
solar thermochemical fuel production. Energy Fuels 26 (3), 19281936. electronic conductivity of Sr1aCeaFe1bCobO3x perovskite-type oxides. J. Eur.
Scheffe, J.R., Steinfeld, A., 2014. Oxygen exchange materials for solar thermochemical Ceram. Soc. 20 (9), 12411250.
splitting of H2O and CO2: a review. Mater. Today 17 (7), 341348. Uhde, 1997. The Shell Gasification Process.
Scheffe, J.R., Allendorf, M.D., Coker, E.N., Jacobs, B.W., McDaniel, A.H., Weimer, A.W., Vandenbossche, M., Mcintosh, S., 2008. The rate and selectivity of methane
2011. Hydrogen production via chemical looping redox cycles using atomic oxidation over La0.75Sr0.25CrxMn1xO3d as a function of lattice oxygen
layer deposition-synthesized iron oxide and cobalt ferrites. Chem. Mater. 23 (8), stoichiometry under solid oxide fuel cell anode conditions. J. Catal. 255 (2),
20302038. 313323.
Scheffe, J.R., Jacot, R., Patzke, G.R., Steinfeld, A., 2013a. Synthesis, characterization, Venstrom, L.J., Petkovich, N., Rudisill, S., Stein, A., Davidson, J.H., 2012. The effects of
and thermochemical redox performance of Hf4+, Zr4+, and Sc3+ doped ceria for morphology on the oxidation of ceria by water and carbon dioxide. J. Sol. Energy
splitting CO2. J. Phys. Chem. C 117 (46), 2410424114. Eng. 134 (1), 11005.
Scheffe, J.R., Weibel, D., Steinfeld, A., 2013b. Lanthanumstrontiummanganese Venstrom, L.J., De Smith, R.M., Hao, Y., Haile, S.M., Davidson, J.H., 2014. Efficient
perovskites as redox materials for solar thermochemical splitting of H2O and splitting of CO2 in an isothermal redox cycle based on ceria. Energy Fuels 28 (4),
CO2. Energy Fuels 27 (8), 42504257. 27322742.
Scheffe, J.R., Welte, M., Steinfeld, A., 2014. Thermal reduction of ceria within an Voitic, G., Hacker, V., 2016. Recent advancements in chemical looping water
aerosol reactor for H2O and CO2 splitting. Ind. Eng. Chem. Res. 53 (6), 2175 splitting for the production of hydrogen. RSC Adv. 6 (100), 9826798296.
2182. Warren, K.J., Reim, J., Randhir, K., Greek, B., Carrillo, R., Hahn, D.W., Scheffe, J.R.,
Schunk, L.O., Lipinski, W., Steinfeld, A., 2009. Heat transfer model of a solar receiver- 2017. Theoretical and experimental investigation of solar methane reforming
reactor for the thermal dissociation of ZnOexperimental validation at 10 kW through the nonstoichiometric ceria redox cycle. Energy Technol.
and scale-up to 1 MW. Chem. Eng. J. 150 (23), 502508. Wei, Y., Wang, H., He, F., Ao, X., Zhang, C., 2007. CeO2 as the oxygen carrier for
Shafiefarhood, A., Hamill, J.C., Neal, L.M., Li, F., 2015. Methane partial oxidation partial oxidation of methane to synthesis gas in molten salts: thermodynamic
using FeOx@La0.8Sr0.2FeO3d coreshell catalyst transient pulse studies. Phys. analysis and experimental investigation. J. Nat. Gas Chem. 16 (1), 611.
Chem. Chem. Phys. 17 (46), 3129731307. Wei, H.J., Cao, Y., Ji, W.J., Au, C.T., 2008. Lattice oxygen of La1xSrxMO3 (M = Mn, Ni)
Shimizu, T., Shimizu, K., Kitayama, Y., Kodama, T., 2001. Thermochemical methane and LaMnO3aFb perovskite oxides for the partial oxidation of methane to
reforming using WO3 as an oxidant below 1173 K by a solar furnace simulator. synthesis gas. Catal. Commun. 9 (15), 25092514.
Sol. Energy 71 (5), 315324. Wei, Y., Wang, H., Li, K., 2010. Ce-Fe-O mixed oxide as oxygen carrier for the direct
Sim, A., Cant, N.W., Trimm, D.L., 2010. Ceriazirconia stabilised tungsten oxides for partial oxidation of methane to syngas. J. Rare Earths 28 (4), 560565.
the production of hydrogen by the methanewater redox cycle. Int. J. Hydrogen Wilhelm, D.J., Simbeck, D.R., Karp, A.D., Dickenson, R.L., 2001. Syngas production for
Energy 35 (17), 89538961. gas-to-liquids applications: technologies, issues and outlook. Fuel Process.
Simakov, D.S.A., Wright, M.M., Ahmed, S., Mokheimer, E.M.A., Romn-Leshkov, Y., Technol. 71 (13), 139148.
2015. Solar thermal catalytic reforming of natural gas: a review on chemistry, Wood, D.A., Nwaoha, C., Towler, B.F., 2012. Gas-to-liquids (GTL): a review of an
catalysis and system design. Catal. Sci. Technol. 5 (4), 19912016. industry offering several routes for monetizing natural gas. J. Nat. Gas Sci. Eng.
Snoeck, J., Froment, G.F., 2003. Steam/CO2 reforming of methane. carbon filament 9, 196208.
formation by the boudouard reaction and gasification by CO2, by H2, and by Wright, H.A., Allison, J.D., Jack, D.S., Lewis, G.H., Landis, S.R., 2003. ConocoPhillips
steam: kinetic study. Fuel Energy Abstr. 44 (3), 147. GTL technology: the COPox process as the syngas generator. ACS Div. Fuel
Snoeck, J.-W., Froment, G.F., Fowles, M., 1997a. Filamentous carbon formation and Chem. Prepr. 48 (2), 791792.
gasification: thermodynamics, driving force, nucleation, and steady-state Yahiro, H., Eguchi, Y., Eguchi, K., Arai, H., 1988. Oxygen ion conductivity of the ceria-
growth. J. Catal. 169 (1), 240249. samarium oxide system with fluorite structure. J. Appl. Electrochem. 18 (4),
Snoeck, J.-W., Froment, G.F., Fowles, M., 1997b. Kinetic study of the carbon 527531.
filament formation by methane cracking on a nickel catalyst. J. Catal. 169 (1), Zeng, Y., Tamhankar, S., Ramprasad, N., Fitch, F., Acharya, D., Wolf, R., 2003. A novel
250262. cyclic process for synthesis gas production. Chem. Eng. Sci. 58 (36), 577582.
Steinfeld, A., 1997. High-temperature solar thermochemistry for CO2 mitigation in Zhao, K., He, F., Huang, Z., Zheng, A., Li, H., Zhao, Z., 2014. Three-dimensionally
the extractive metallurgical industry. Energy 22 (23), 311316. ordered macroporous LaFeO3 perovskites for chemical-looping steam reforming
Steinfeld, A., Meier, A., 2004. Solar fuels and materials. Encyclopedia of Energy. of methane. Int. J. Hydrogen Energy 39 (7), 32433252.
Elsevier, pp. 623637. Zhao, K., He, F., Huang, Z., Wei, G., Zheng, A., Li, H., Zhao, Z., 2016. Perovskite-type
Steinfeld, A., Spiewak, I., 1998. Economic evaluation of the solar thermal co- oxides LaFe1xCoxO3 for chemical looping steam methane reforming to syngas
production of zinc and synthesis gas. Energy Convers. Manag. 39 (15), 15131518. and hydrogen Co-production. Appl. Energy 168, 193203.
Steinfeld, a., Kuhn, P., Karni, J., 1993. High-temperature solar thermochemistry: Zheng, Y., Zhu, X., Wang, H., Li, K., Wang, Y., Wei, Y., 2014a. Characteristic of
production of iron and synthesis gas by Fe3O4-reduction with methane. Energy macroporous CeO2-ZrO2 oxygen carrier for chemical-looping steam methane
18 (3), 239249. reforming. J. Rare Earths 32 (9), 842848.
Steinfeld, A., Frei, A., Kuhn, P., 1995a. Thermoanalysis of the combined Fe3O4- Zheng, Y., Wei, Y., Li, K., Zhu, X., Wang, H., Wang, Y., 2014b. Chemical-looping steam
reduction and CH4-reforming processes. Metall. Mater. Trans. B 26 (3), 509515. methane reforming over macroporous CeO2ZrO2 solid solution: effect of
Steinfeld, A., Frei, A., Kuhn, P., Wuillemin, D., 1995b. Solar thermal production of calcination temperature. Int. J. Hydrogen Energy 39 (25), 1336113368.
zinc and syngas via combined ZnO-reduction and CH4-reforming processes. Int. Zhou, G., Shah, P.R., Kim, T., Fornasiero, P., Gorte, R.J., 2007. Oxidation entropies and
J. Hydrogen Energy 20 (10), 793804. enthalpies of ceria-zirconia solid solutions. Catal. Today 123, 8693.
Steinfeld, A., Larson, C., Palumbo, R., Foley, M., 1996. Thermodynamic analysis of the Zhu, X., Wang, H., Wei, Y., Li, K., Cheng, X., 2010. Hydrogen and syngas production
co-production of zinc and synthesis gas using solar process heat. Energy 21 (3), from two-step steam reforming of methane over CeO2-Fe2O3 oxygen carrier. J.
205222. Rare Earths 28 (6), 907913.
Steinfeld, A., Brack, M., Meier, A., Weidenkaff, A., Wuillemin, D., 1998. A solar Zhu, X., Wang, H., Wei, Y., Li, K., Cheng, X., 2011a. Reaction characteristics of
chemical reactor for Co-production of zinc and synthesis gas. Energy 23 (10), chemical-looping steam methane reforming over a CeZrO2 solid solution
803814. oxygen carrier. Mendeleev Commun. 21 (4), 221223.
Stobbe, E.R., de Boer, B.A., Geus, J.W., 1999. The reduction and oxidation behaviour Zhu, X., Wang, H., Wei, Y., Li, K., Cheng, X., 2011b. Hydrogen and syngas production
of manganese oxides. Catal. Today 47 (14), 161167. from two-step steam reforming of methane using CeO2 as oxygen carrier. J. Nat.
Suzuki, T., Nakayama, O., Okamoto, N., 2012. Partial oxidation of methane to Gas Chem. 20 (3), 281286.
nitrogen free synthesis gas using air as oxidant. Catal. Surv. from Asia 16 (2), Zhu, X., Wei, Y., Wang, H., Li, K., 2013. CeFe oxygen carriers for chemical-looping
7590. steam methane reforming. Int. J. Hydrogen Energy 38 (11), 44924501.
Takacs, M., Scheffe, J.R., Steinfeld, A., 2015. Oxygen nonstoichiometry and Zhu, X., Li, K., Wei, Y., Wang, H., Sun, L., 2014. Chemical-looping steam methane
thermodynamic characterization of Zr doped ceria in the 15731773 K reforming over a CeO2Fe2O3 oxygen carrier: evolution of its structure and
temperature range. Phys. Chem. Chem. Phys. 17 (12), 78137822. reducibility. Energy Fuels 28 (2), 754760.
Takenaka, S., Hanaizumi, N., Son, V.T.D., Otsuka, K., 2004. Production of pure Zinkevich, M., Djurovic, D., Aldinger, F., 2006. Thermodynamic modelling of the
hydrogen from methane mediated by the redox of Ni- and Cr-added iron oxides. ceriumoxygen system. Solid State Ionics 177 (1112), 9891001.
J. Catal. 228 (2), 405416.

Please cite this article in press as: Krenzke, P.T., et al. Solar fuels via chemical-looping reforming. Sol. Energy (2017), http://dx.doi.org/10.1016/j.
solener.2017.05.095

You might also like