You are on page 1of 11

Solar Energy 174 (2018) 617–627

Contents lists available at ScienceDirect

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

Solar electricity via an Air Brayton cycle with an integrated two-step T


thermochemical cycle for heat storage based on Fe2O3/Fe3O4 redox
reactions: Thermodynamic and kinetic analyses
H. Evan Bush, Peter G. Loutzenhiser

George W. Woodruff School of Mechanical Engineering, Georgia Institute of Technology, Atlanta, GA 30332-0405, USA

ARTICLE INFO ABSTRACT

Keywords: Solar electricity production via an Air Brayton cycle is considered with integrated thermochemical energy
Concentrated solar energy storage. The storage is realized via a two-step solar thermochemical cycle based on Fe2O3/Fe3O4 reduction-
Thermochemical energy storage oxidation reactions, encompassing (1) the thermal reduction of Fe2O3 to Fe3O4 and O2 driven by concentrated
Iron oxide solar irradiation under vacuum; and (2) the exothermic oxidation of Fe3O4 with a compressed air stream back to
Thermodynamic analysis
Fe2O3. The steps may be decoupled, resulting in a high temperature, pressurized airflow that is expanded across
Kinetic analysis
a turbine to produce on-demand electricity. A thermodynamic analysis of the system determined a maximum
cycle efficiency of 46.0% at a solar concentration ratio of 4000 suns, an oxidation pressure of 30 bar, and an
approximately 5:1 molar flow rate ratio of air to solid Fe2O3 exiting the re-oxidizer. Chemical kinetics for the
thermal reduction of Fe2O3 were determined between approximately 1400 and 1700 K using non-isothermal
thermogravimetry with heating rates between 10 and 20 K·s−1 and O2 partial pressures between 0 and 0.05 bar.
The rate-limiting reaction mechanism was determined to be nucleation, and kinetic parameters were resolved
using an Avrami-Erofe’ev nucleation model with a reaction order of 1.264 ± 0.010. The rate constant followed
an Arrhenius-type temperature dependency with an apparent activation energy of 487.0 ± 3.6 kJ·mol−1 and
pre-exponential factor 2.768 ± 0.783·1014 s−1. A power-law dependence on O2 partial pressure of order
8.317 ± 0.233 was determined. Non-isothermal thermogravimetry to examine the oxidation of Fe3O4 to Fe2O3
revealed multiple kinetic regimes, and isothermal thermogravimetry showed the reaction proceeded rapidly,
within 20 s, at temperatures greater than 673 K. Solid characterization was carried out using scanning electron
microscopy and x-ray powder diffractometry up to temperatures of 1073 K to verify initial and final sample
compositions and structures.

1. Introduction efficiency and producing a mismatch in pairing to solar concentrating


infrastructure capable of achieving higher solar concentration ratios
In previous works, thermodynamic and kinetic analyses were used (Steinfeld and Palumbo, 2001).
to examine electricity production in an Air Brayton cycle with in- The purpose of this work is to thermodynamically and kinetically
tegrated solar thermochemical energy storage (TCES) based on Co3O4/ examine a similar cycle based on the Fe2O3/Fe3O4 redox pair. Fe2O3/
CoO oxidation-reduction (redox) reactions at a solar concentration ratio Fe3O4 is one of many metal oxide candidates proposed for TCES ap-
of 1000 suns (where 1 sun = 1 kW·m−2) (Muroyama et al., 2015; plications, including the Co3O4/CoO (Agrafiotis et al., 2014; Hutchings
Schrader et al., 2017; Schrader et al., 2015). TCES based on the Co3O4/ et al., 2006; Neises et al., 2012; Pagkoura et al., 2014), Mn2O3/Mn3O4
CoO redox pair is especially promising due to rapid achievable reaction (Alonso et al., 2013; Carrillo et al., 2014), CuO/Cu2O (Alonso et al.,
rates, a high reaction enthalpy, and cyclability of the reversible oxi- 2015; Deutsch et al., 2017; Haseli et al., 2017), and BaO2/BaO (Bowrey
dation/thermal reduction reactions. However, cobalt oxides are rela- and Jutsen, 1978; Carrillo et al., 2016) binary pairs, as well as a number
tively sparse and expensive (Wong, 2011) and pose potential human of mixed ionic-electronic conducting metal oxides, which have the
and environmental health concerns (SIGMA-ALDRICH, 2017). The advantages of continuous redox equilibria, tunable thermodynamic
Co3O4/CoO redox pair also thermally reduces at a lower temperature properties, high reaction rates, and large redox capacities without de-
compared to other redox pairs, reducing the maximum achievable cycle parture from the perovskite phase (Babiniec et al., 2015, 2016;


Corresponding author.
E-mail address: peter.loutzenhiser@me.gatech.edu (P.G. Loutzenhiser).

https://doi.org/10.1016/j.solener.2018.09.043
Received 14 July 2018; Received in revised form 14 September 2018; Accepted 16 September 2018
Available online 24 September 2018
0038-092X/ © 2018 Elsevier Ltd. All rights reserved.
H.E. Bush, P.G. Loutzenhiser Solar Energy 174 (2018) 617–627

Nomenclature 0 environment/dead state (298.15 K and 1 bar), initial value


air air, approximated as a 21% O2–N2 concentration ideal gas
ΔG Gibbs property/free enthalpy mixture
ΔH enthalpy of reaction comp cycle air compressor (for re-oxidizer) stage
Δm TGA mass loss/gain cool receiver outlet O2 cooling/quench stage
A undetermined constant cycle combined thermochemical and Air Brayton cycle
C solar concentration ratio (relative to 1 kW·m−2) eq thermodynamic equilibrium
D maximum particle size for oxidation of magnetite to ma- exhaust air stream exiting turbine
ghemite g gas state
Ea apparent activation energy i index variable
f model equation/kinetic solid conversion dependence j index variable
(differential form) loss re-radiative and convective thermal losses
h kinetic O2 partial pressure dependence m maximum
k rate constant/kinetic temperature dependence oct octahedrally-configured
k0 apparent pre-exponential/frequency factor reactor cycle solar receiver/reactor (reduction) stage
m mass, partial pressure-dependence rate constant recover Fe2O3/Fe3O4 particle stream exiting re-oxidizer
n Avrami-Erofe’ev rate constant solar cycle solar thermochemical reactor stage
n molar flow rate tet tetrahedrally-configured
p pressure, unknown model parameter turbine Air Brayton cycle turbine stage
Q heat transfer rate vac cycle vacuum pump (for solar receiver/reactor) stage
r reaction/temporal conversion rate
r2 coefficient of determination Superscripts/accents
R universal gas constant
SSE sum of squared errors/residual sum of squares ^ dependent variable, model estimate
t time
T temperature Acronyms
W power
x nonstoichiometry AN nth-order Avrami-Erofe’ev nucleation-limited model
FCC face-centered cubic or cubic close-packed crystalline
Greek letters structure
HCP hexagonal close-packed crystalline structure
α conversion, hematite SEM scanning electron microscopy
β TGA heating rate TCES thermochemical energy storage
γ maghemite TGA thermogravimetric analysis/analyzer
η efficiency XRD X-ray diffractometry

Subscripts

∞ equilibrium value

Imponenti et al., 2017; Zhang et al., 2016). Fe2O3/Fe3O4 redox reac- metals, with the goal of lowering the thermal reduction temperature
tions are particularly attractive because Fe2O3 has a higher thermal (André et al., 2017; Block et al., 2014; Block and Schmücker, 2016;
reduction temperature than most other binary candidates, permitting Pagkoura et al., 2014). However, past efforts have shown a significant
higher theoretical Air Brayton cycle efficiencies. The Fe2O3/Fe3O4 decrease in reaction enthalpy.
redox pair materials are also are relatively inexpensive and widely In this study, the Fe2O3 thermal reduction temperature was reduced
available compared to Co3O4/CoO materials, and they carry fewer en- by operating at decreased O2 partial pressures to shift the equilibrium
vironmental/human health concerns (Smith and Huyck, 1999), making to a more favorable temperature range according to Le Chatelier’s
the redox pair a more practical and economically appealing candidate principle. The impact of decreasing the O2 partial pressure is shown in
for large-scale deployment. The reversible redox reaction of Fe2O3/ Fig. 1. For an O2 partial pressure of 10−3 bar, Fe2O3 begins to reduce to
Fe3O4 is represented as: Fe3O4 at a temperature of 1432 K. The Fe2O3/Fe3O4 pair allows for
higher oxidation temperatures and, therefore, greater Air Brayton cycle
1
3Fe2 O3 2Fe3O4 + O2(g) , H298.15 K = 78.3 kJ·mol 1
theoretical efficiencies, at the cost of decreasing receiver absorption
2 (1)
efficiencies due to greater re-radiative losses to the environment
Iron oxides have been the focus of numerous solar thermochemical (Steinfeld and Palumbo, 2001). Therefore, a comprehensive thermo-
studies, including carbothermal iron production (Steinfeld and Fletcher, dynamic cycle analysis was performed to assess the potential of the
1991), fuels co-production via carbothermal reduction and/or methane Fe2O3/Fe3O4 pair.
cracking (Halmann and Steinfeld, 2006; Steinfeld et al., 1995; Steinfeld An analysis of Fe2O3/Fe3O4 TCES integrated into an Air Brayton
et al., 1993; Tamaura et al., 1997), and H2O and/or CO2 splitting via cycle was performed as a function of molar flow rate of air, reactor
two-step cycles based on Fe3O4/FeO redox reactions (Galvez et al., temperature, compressor outlet/turbine inlet pressure, and solar con-
2008; Loutzenhiser et al., 2009; Nakamura, 1977; Sibieude et al., 1982; centration ratio. The thermal reduction and oxidation reactions were
Stamatiou et al., 2010). For the Fe2O3/Fe3O4 pair, the high Fe2O3 examined using thermogravimetric analysis for a range of heating rates
thermal reduction temperatures have presented both materials com- and O2 partial pressures to measure reaction rates, identify rate limiting
patibility and engineering design challenges. Such concerns have mechanism(s), and determine kinetic parameters. A cycling study was
prompted the study of combined oxides of iron and other multivalent performed to examine material stability and scanning electron

618
H.E. Bush, P.G. Loutzenhiser Solar Energy 174 (2018) 617–627

were extracted from the NIST Chemistry Webbook (Linstrom and


Mallard, 2001). The thermal reduction step was fixed at constant values
ranging from Treactor = TΔG=0 = 1432 K, corresponding to ΔG = 0 for
an oxygen partial pressure pO2 =10−3 bar (i.e., conditions where the
reaction becomes spontaneous) to Treactor = 1632 K, selected due to
material limitations and rapidly increasing radiative losses. The solar
thermochemical reactor was idealized as a windowed cavity comprised
of two diffuse, isothermal surfaces, one opaque and one semi-trans-
parent to shortwave radiation, and radiative exchange was modeled
using a two-band gray approximation. Full conversion of Fe2O3 to
Fe3O4 and O2 at solar concentration ratios between C = 1000 and 5000
suns was assumed due to the thermodynamic favorability of the thermal
reduction that resulted from the continuous removal of O2 by the va-
cuum pump during the process. The analysis was normalized to 2/
3 mol·s−1 of Fe3O4 exiting the solar thermochemical reactor.
The re-oxidizer was modeled in a counterflow configuration with
the minimum particle outlet temperature Trecover limited to the tem-
perature of air exiting the compressor Tcomp. Airflow into the com-
pressor was controlled to ensure Tturbine, the temperature of O2-defi-
Fig. 1. Change in Gibb’s free energy as a function of temperature for the re- cient air exiting the re-oxidizer and entering the turbine, was not so
duction of Fe2O3 at O2 partial pressures of 0.001, 1, and 6 bar. high as to prevent re-oxidation. In order to define an upper bound on
cycle performance independent of subcomponent designs, the com-
pressor and turbine were idealized as operating isentropically, and the
microscopy and room and high-temperature x-ray diffractometry were vacuum pump used to hold the reactor at a low vacuum pressure preactor
performed to study particle size, morphology, post-reaction chemical as isothermal and reversible at the environment temperature
composition, and Fe2O3/Fe3O4 crystalline structure. T0 = 298.15 K. The work, heat, and pressure losses from material
transport/storage were neglected as in previous analyses (Schrader
2. Thermodynamic analysis et al., 2015). Rough estimates of these nonidealities for similar ther-
modynamic cycles have been made elsewhere (Gorman et al., 2015), as
A schematic from previous work (Schrader et al., 2015), modified have estimates for the work requirements of real vacuum pumps and/or
for Fe2O3/Fe3O4 redox reactions, is provided in Fig. 2. The cycle em- air separation methods, both of which are expected be significantly
ploys a windowed, directly irradiated solar receiver/reactor at low higher than ideal performance calculations (Brendelberger et al., 2017).
vacuum ( pO2 = 10−3 bar) to promote the thermal reduction of metal The Trecover and Tturbine resulting from the re-oxidizer chemical
oxide particles at a temperature Treactor. Reduced particles are then re- equilibrium and energy balances for fixed Treactor = 1432 K and
oxidized in a counter-flow configuration with a pressurized airflow at pcomp = 30 bar are shown in Fig. 3. For the entire range of nair shown,
pcomp supplied by a compressor. The high-temperature, O2-deficient air the entering Fe3O4 particle flow was fully re-oxidized to Fe2O3. Tturbine
exits the re-oxidizer and expands through a turbine, while the re-oxi- gradually increased until nair = 5.1 mol·s−1, for which the sensible heat
dized particles are returned to the receiver/reactor to complete the extracted from the solid particle flow was maximized due to Trecover
cycle. reaching the minimum thermodynamically allowable value of
A thermodynamic analysis of the system was performed in- Trecover = Tcomp. For nair >5.1 mol·s−1, Tturbine therefore rapidly de-
dependently of subcomponent design considerations and kinetic lim- creased.
itations to ascertain theoretical efficiency limits by imposing first and Fig. 4 shows the resulting subcomponent heat transfer rate Q and
second thermodynamic law constraints. Ideal gas behavior, homo- power W values for the same Treactor and pcomp as Fig. 3 and C = 1000
geneous gas/solid mixtures, and local thermal equilibrium between the suns. The heat rejection via O2 exiting the reactor Qcool remained con-
gas and solid phases were assumed, and thermophysical properties stant, as complete re-oxidation occurred for all nair shown. Consistent

Fig. 2. System schematic of the Air Brayton cycle with an integrated two-step thermochemical energy storage cycle based on Fe2O3/Fe3O4 redox reactions with
relevant mass and energy flows.

619
H.E. Bush, P.G. Loutzenhiser Solar Energy 174 (2018) 617–627

Table 1
Model property values for the analysis of cycle efficiency vs. solar concentration
ratio for the Air Brayton cycle with paired thermochemical energy storage using
Fe2O3/Fe3O4 and Co3O4/CoO reduction-oxidation pairs.
Property Fe2O3/Fe3O4 Co3O4/CoO

Treactor 1037 K 1432 K


nair 5.1 mol·s−1 13.6 mol·s−1
pcomp 30 bar 30 bar
pO 2 0.001 bar 0.001 bar

Fig. 3. Fe2O3/Fe3O4 particle temperature exiting the re-oxidizer (solid) and O2-
depleted air temperature entering the turbine (dashed), as functions of molar
air flow rate and for a solar thermochemical reactor temperature of 1432 K,
solar concentration ratio of 1000 suns, and re-oxidizer pressure of 30 bar.

with the abrupt Trecover minimum in Fig. 3, the required input reactor
heat transfer rate Qsolar reached a maximum for nair=5.1 mol·s−1, as the
particles were fully re-oxidized at the minimum allowable temperature,
providing no more chemical or sensible energy storage potential for
higher nair . The turbine power Wturbine = increased slower for
nair >5.1 mol·s−1 but did not become zero, as the increasing nair con-
Fig. 5. Theoretical cycle efficiencies for Fe2O3/Fe3O4 (solid) and Co3O4/CoO
tinued to increase the power output.
(dashed) thermochemical cycles as a function of solar concentration ratio at
The maximum theoretical thermodynamic cycle efficiency for the receiver temperatures of 1432 and 1037 K and compressor airflow rates of 5.1
Air Brayton-integrated thermochemical cycle was defined as: and 13.6 mol·s−1, respectively, and a re-oxidizer pressure of 30 bar.
Wturbine Wcomp Wvac
=
(2)
cycle
Qsolar
Table 1.
At C = 1000 suns, Treactor = 1432 K, pcomp = 30 bar, and The ηcycle for Fe2O3/Fe3O4 increased more rapidly with C than for
nair = 5.1 mol·s−1, a ηcycle = 38.4% was determined, which is sig- Co3O4/CoO due to a higher Treactor, as shown in Fig. 5, and the ηcycle for
nificantly lower than the ηcycle = 43.4% from the Co3O4/CoO analysis the Fe2O3/Fe3O4 TCES cycle exceeded the Co3O4/CoO cycle at
under equivalent conditions (Schrader et al., 2015). The lower ηcycle C ≥ 4352 suns.
was primarily due to significantly higher Treactor and, therefore, higher Fig. 6 shows nair for a range of pcomp and Treactor. A ηcycle = 46.0%
Qloss for the Fe2O3/Fe3O4 pair. However, because Qloss is mitigated by was achieved for C = 4000 suns, Treactor = 1432 K, pcomp = 30 bar, and
increasing C (Steinfeld et al., 1998; Steinfeld and Palumbo, 2001), an nair = 5.1 mol·s−1, and the ηcycle increased with diminishing returns for
analysis of both redox pairs was performed to determine how ηcycle greater nair and C. Increasing pcomp increased ηcycle as the O2-deficient
varies with C for Treactor = TΔG=0 and nair fixed to achieve full re-oxi- airflow entering the turbine expanded across a greater temperature
dation and maximum sensible energy recovery in the re-oxidizer. Re- difference, while increasing Treactor decreased ηcycle due to rapid in-
levant model property value inputs for the analysis are shown in creases in Qloss from reradiation.

Fig. 4. Energy balance components for the in-


tegrated thermochemical cycle, including (a) for
the solar thermochemical reactor, heat transfer
rates associated with concentrated solar irradia-
tion (solid), thermal losses (dashed), and re-
jected high-temperature O2 from thermal re-
duction (dashed-dot); and (b) power associated
with turbine output (solid), compressor input
(dashed), and vacuum pump input (dashed-dot)
as functions of the molar air flow rate and for a
solar thermochemical reactor temperature of
1432 K, solar concentration ratio of 1000 suns,
and a re-oxidizer pressure of 30 bar.

620
H.E. Bush, P.G. Loutzenhiser Solar Energy 174 (2018) 617–627

Fig. 6. Theoretical cycle efficiency as a function of molar air flow rate entering the re-oxidizer for (a) compressor pressures of 5, 17.5, and 30 bar and a solar
thermochemical reactor temperature of 1432 K; and (b) solar thermochemical reactor temperatures of 1432, 1532, and 1632 K and a compressor pressure of 30 bar.

3. Kinetic analyses confidence intervals for the parameters p1 … pj using the 2.5th and
97.5th percentile values of each fitted parameter (Alper and Gelb,
The thermal reduction of Fe2O3 to Fe3O4 and O2 and re-oxidation of 1990).
Fe3O4 to Fe2O3 were examined in a thermogravimetric analyzer (TGA,
Netzsch STA 449 F3 Jupiter, ± 1 μg). Separate unreacted, high-purity
Fe2O3 and Fe3O4 samples were selected for the two reaction steps to 3.1. Fe2O3 thermal reduction
ensure that samples were not initially in partially reduced or oxidized
states, respectively. Powder samples were placed on an Al2O3 crucible The rate limiting mechanism of the Fe2O3 thermal reduction was
shielded with a platinum foil (Sigma-Aldrich 0.025MM Thick 99% Pt described with a variable-order Avrami-Erofe’ev nucleation-controlled
Foil 267244-1.4G). Sample temperatures were measured by an S-type model (AN), given as:
thermocouple. Sample masses were measured before and after testing
n-1
with an analytical balance (Mettler-Toledo ML54, ± 0.1 mg) to verify f ( ) = n (1 )[ ln(1 )] n (7)
the total mass change measured by TGA. Blank runs under identical
conditions were performed following each experiment to correct for the where n is the reaction order. k0, Ea, and n were simultaneously fitted at
influences of buoyancy and gas dynamics. multiple heating rates to verify that the reaction was not heat transfer
The temporal conversion from Fe2O3 to Fe3O4 and vice versa were limited (Vyazovkin et al., 2011). The three fitted values were then fixed
calculated as: and h (pO2 ) was described using a power-law relationship, given as:

m (t ) m 0 pO2
m
(t ) =
m m0 (3) h (pO2 ) = 1
pO2, eq (T ) (8)
where m(t) is the mass at time t; m0 is the initial sample mass, between
90 and 100 mg; and m∞ is the stoichiometric equilibrium mass for the where pO2,eq is the equilibrium partial pressure for T and m is the
fully-reacted sample. The reaction rate was defined as: constant reaction order for the pO2 dependence, which was simulta-
neously fitted to experimental runs performed at multiple pO2 .
d
r= = h (pO2 ) k (T ) f ( ) All non-isothermal reduction experiments were performed with a
dt (4)
constant total gas flow of 100 mLN·min−1 (where LN refers to liters at
where the temporal conversion rate dα/dt was calculated with a fourth- standard conditions, i.e., 273 K and 1 bar). Fe2O3 samples (Sigma-
order finite difference approximation; and a reaction rate constant k(T), Aldrich ≥99.995 trace metals basis, particle size < 5 μm) were heated
partial pressure dependence h (pO2 ) , and conversion-dependent kinetic to 1723 K in Ar at heating rates of 10–20 K·min−1 in the TGA high
model f(α) were fitted to r using a Nelder-Mead nonlinear minimization temperature graphite furnace. The thermal reduction proceeded at
algorithm with a standard squared error objective function, given as: elevated temperatures, and samples agglomerated and turned dark
gray, similar to unreacted Fe3O4.
SSE = [ri ri (t , T, , pO2 ; p1, p2, …, pj)]2 The relationship between α and T was found to be independent of β,
i (5)
indicating no heating rate (heat transfer) limitations. The reaction rate
where a least squares criterion was used to determine r , the modeled consistently slowed significantly for values α > 0.90, and an iso-
reaction rate; t, T, α, and pO2 were known for the datapoints i; and the conversional analysis (Friedman, 1964) revealed large changes in Ea for
unknown parameters p1 … pj were simultaneously fitted. An Arrhenius- such values. Thus, the analysis was limited to 0 < α < 0.90, and
type temperature dependency was assumed, given as: k0 = 2.768 ± 0.783 · 1014 s−1, Ea = 487.0 ± 3.6 kJ·mol−1, and
n = 1.264 ± 0.010 were simultaneously fitted with 95% confidence
Ea
k (T ) = k 0 exp intervals from the Monte Carlo analysis.
RT (6)
Two methodologies were employed to validate the model-fitted
where k0 is the apparent pre-exponential factor; Ea is the apparent ac- parameters. To evaluate the nucleation model, reaction order, and the
tivation energy; and R is the universal gas constant. A Monte Carlo error appropriateness of the Arrhenius form of k, an Arrhenius plot was
analysis using 500 iterations was performed to determine 95% produced by calculating k as:

621
H.E. Bush, P.G. Loutzenhiser Solar Energy 174 (2018) 617–627

temperatures. The model accurately captured changes in α as a function


of pO2 for pO2 < 0.05 bar, a range which could reasonably be attained
with a windowed receiver, vacuum pump and/or inert gas flow.

3.2. Fe3O4 oxidation

Fe3O4 oxidation to Fe2O3 has been described by a series of structure-


dependent intermediate reactions detailed in previous work (Feitknecht
and Gallagher, 1970; Gallagher et al., 1968; Shebanova and Lazor,
2003; Sahoo et al., 2010). Fe3O4 has a cubic [Fe2+][Fe3+]2O inverse
spinel structure with O2− anions ordered in a close-packed face-cen-
tered cubic (FCC) lattice and Fe cations that occupy interstitial loca-
tions within the lattice: Fe2+ cations at tetrahedral sites Fetet and Fe3+
cations evenly divided between octahedral Feoct and tetrahedral sites
(Parkinson, 2016). The remaining interstitial sites exist as cation va-
cancies, at a 1:1 vacancy to cation ratio for octahedral sites and 7:1
ratio for tetrahedral sites.
As shown in Fig. 10a, oxidation initially proceeds via Feoct ion dif-
fusion through the particle to bond with adsorbed O2−, accompanied
by Fe2+→Fe3+ oxidation, which results in the formation of cation va-
Fig. 7. An Arrhenius plot of the reaction rate constant for the non-isothermal
thermal reduction of Fe2O3 at 10–20 K·min−1, calculated from: (1) the reaction
cancies VFe (Becker et al., 1993; Sidhu et al., 1977). The process is re-
model, fitted reaction order, and measured reaction rates (markers), and (2) the presented as:
fitted pre-exponential factor and activation energy (line). 2x 3
Fe3 O4 + O2(g) Fe3 - x O4
3 x 3 x (11)
ri where x is the deviation from stoichiometry. As shown in Fig. 10b, in
k (Ti ) =
f ( i, n ) (9) sufficiently small particles of diameter d p < D=300 nm (Feitknecht and
Gallagher, 1970), and under thermodynamically favorable conditions,
where ri, αi, and Ti are the same experimental data used to compute the
the reaction proceeds until xm = 1/3 and the Fe3+:O2− ratio is 2:3,
nonlinear fit for k0, Ea and n. Fig. 7 contains the Arrhenius plot with
while the anion lattice structure remains unchanged, resulting in cubic
calculated (markers) k’s and modeled (line) values as a function of
-Fe2O3, or maghemite. Elevated temperatures induce nucleation of
temperature, where the modeled k curve was determined from the
rhombohedral α-Fe2O3, or hematite, in which the O2− anion lattice
fitted k0 and Ea above. The calculated k’s showed a clear Arrhenius
transitions to a hexagonal close packed (HCP) structure and the Fe3+
temperature relationship for the fitted reaction order. The calculated
cations occupy octahedral interstitial positions.
results were matched by the modeled curve for most temperatures,
As shown in Fig. 10c, when d p > D , a phase change occurs at
although the curve somewhat underpredicted k at low temperatures
x m < 1 3 with sufficiently high temperatures and cation vacancy gra-
and β’s (i.e., where the fixed measurement resolution and steeper ri
dients to induce lattice strain, resulting in a disproportionated solid
during early conversion would increase numerical errors).
solution of α-Fe2O3 and stoichiometric Fe3O4, represented as:
Ea was also calculated independently of the conversion model from
the same experiments using an Avrami plot and the Kissinger method Fe3 - xm O4 [4x m ]Fe2 O3 + [1 3x m ]Fe3 O4 (12)
(Kissinger, 1957), given as:
where x m may vary with β and temperature due to their influence on
Ea 1
ln = +A
Tm2 R Tm (10)

where Tm is the temperature at which r was maximized in each ex-


periment, A is a constant, and Ea was determined from linear regression.
Fig. 8 is an Avrami plot and accompanying linear regression applied
with the Kissinger method to determine Ea = 574 kJ·mol−1.
Previous analyses of the reduction of Fe2O3 in CO, H2, or reducing
H2O–H2 environments used both AN nucleation and zero-order phase
boundary reaction mechanisms, and found that reduction occurred at
lower temperatures and with a lower range of
Ea = 33.28–139.2 kJ·mol−1 (Tiernan et al., 2001; Yu et al., 2017).
While varying factors such as sample purity, specific surface area, and
particle size affect activation energy, the much higher fitted
Ea = 487.0 kJ·mol−1 for an inert environment was consistent with the
higher reduction temperatures and suggested a gas composition de-
pendence.
Further non-isothermal thermal reduction experiments were per-
formed for 2–5% O2–Ar to evaluate the impact of pO2 with the de-
termined kinetic parameters, and m = 8.317 ± 0.233 was determined
via non-linear regression. The modeled and measured α’s are plotted as
Fig. 8. An Avrami plot and regression line for the non-isothermal thermal re-
a function of temperature in Fig. 9. The temperature of the onset of
duction of Fe2O3 at 10–20 K·min−1, where the five markers correspond to the
reduction rose for increasing pO2 due to thermodynamic limitations as temperature of maximum conversion rate during an experiment at 10, 12.5, 15,
descibed by Le Chatelier’s Principle, while the temperature range for 17.5 and 20 K·min−1 respectively.
,
the reaction was significantly reduced due to an increased k at higher

622
H.E. Bush, P.G. Loutzenhiser Solar Energy 174 (2018) 617–627

Fig. 11. Experimental conversion (markers) and temporal conversion rate


(solid line) for non-isothermal Fe3O4 oxidation at O2 concentrations between
20% and 80%.

closer to conditions expected in the re-oxidizer. Samples were heated in


100% Ar at β = 20 K·min−1 to between T = 673 and 973 K, followed by
a 20 min isothermal step to allow the sample to equilibrate. The flow
Fig. 9. Measured (markers) and predicted (lines) conversions of Fe2O3 to Fe3O4
was then switched to 80% O2-Ar, after which the oxidation reaction
as functions of temperature and O2 concentration between 0 and 5% O2–Ar.
proceeded rapidly (within t < 20 s) but did not achieve complete
conversion, as shown in Fig. 12a-b. Compared to the non-isothermal
cation vacancy gradients and lattice strain. As shown in Fig. 10d, fur- experiments, the first peak in d dt occurred at lower 0.12 , re-
ther oxidation of the remaining Fe3O4 occurs after disproportionation sulting in an abrupt change in the slope of the 's in Fig. 12a. For
rather than via the nonstoichiometric reaction. Oxidation then proceeds > 0.12, r was mostly invariant between experiments. The total oxi-
via direct conversion to α-Fe2O3 facilitated by α-Fe2O3 nucleation sites, dation time differed almost solely due to varying r for 0.12, as
represented as: shown in Fig. 12b. As D < dp < 88 μm for the Fe3O4 particles, the
earlier α transition in isothermal experiments may have resulted from
1 3x m 1 3x m 3 xm
Fe2 O3 + Fe3O4 + O2(g) Fe2 O3 an effective minimization of x m , and therefore earlier onset of α-Fe2O3
4x m 16x m 8x m (13)
nucleation, by initiating oxidation at higher temperatures.
where together, Eqs. (11)–(13) are the back reaction of Equation (1). Two step models have been previously used to capture Fe3O4 oxi-
The oxidation of Fe3O4 to Fe2O3 was examined via TGA with a dation. A study of Fe3O4 pellet oxidation in air found two oxidation
constant total gas flow of 100 mLN·min−1. Fe3O4 particle samples (Alfa regimes: (1) for T 693 K, a surface-controlled regime with
Aesar Puratronic® 99.997% metals basis) were sieved with a size #170 Ea ≈ 155 kJ·mol−1, and (2) for T 693 K, a diffusion-controlled regime
(88 μm) mesh before each experiment. Samples were heated to 1623 K with 10.5 kJ·mol−1 (Papanastassiou and Bitsianes, 1973). In more re-
in 80% O2-Ar in the TGA high temperature graphite furnace at cent work, Fe3O4 pellet oxidation in air for 1023–1173 K was described
10–20 K·min−1, where different β’s were used to ensure the reaction by (1) a nucleation mechanism, with an Ea = 4.21 ± 0.45 kJ·mol−1,
was not heat transfer limited. Additional non-isothermal experiments then (2) a diffusion mechanism, with Ea = 53.58 ± 3.56 kJ·mol−1
were performed for 20%–80% O2-Ar to examine pO2 dependence, as (Monazam et al., 2014). Both studies used pellets of dp 88 μm, which
shown in Fig. 11. As no such dependence was observed, it was assumed presumably accounted for a diffusion-limited regime following nu-
that the reaction was not limited by surface O2 adsorption for cleation. However, in this study, isothermal thermogravimetry was
pO2 > 0.2 bar, including conditions expected in the re-oxidizer. For strongly impacted by both rapid, complex reaction behavior at low α
each experiment, two local maxima in d dt were observed at (1) T ≈ and uncertainty due to transients related to gas switching. As a result,
525 K and (2) T ≈ 650 K, as shown for a single d dt curve in Fig. 11. both the non-isothermal and isothermal methods were inadequate for
This translated to an inflection point at 0.74 , indicated by a vertical modeling Fe3O4 oxidation and determining rate-limiting mechanism(s),
dashed line. The isoconversional method (Friedman, 1964) revealed and this was left for future work.
signficant changes in Ea near the same α. As the proposed oxidation
pathway contains multiple reaction steps with different mechanisms, it 4. Reduction-oxidation cycling
is likely that the second d dt maximum and observed Ea changes re-
sulted due to an apparent two-step oxidation process. Fe2O3 samples were cycled between reduced and oxidized states via
Because oxidation reached high α’s at T < 620K , a suite of iso- an isothermal pO2 swing to evaluate their suitability for use in TCES
thermal experiments was performed to study oxidation at higher T, cycles. Samples were heated to and held at 1623 K under an oxidizing

Fig. 10. Schematic of Fe3O4 oxidation pathways (a) at the onset of the diffusion-limited regime, resulting in (b) oxidation to maghemite, or γ-Fe2O3; and (c) at the
transition to the nucleation-limited regime, resulting in (d) oxidation to hematite, or α-Fe2O3.

623
H.E. Bush, P.G. Loutzenhiser Solar Energy 174 (2018) 617–627

Fig. 12. Experimental conversion for isothermal Fe3O4 oxidation at an O2 concentration of 80% and temperatures between 673 and 973 K: (a) the full oxidation
profiles and (b) detail-views of the first 25 s of oxidation.

environment, then cycled between 0% and 20% O2–Ar in 15 min in- to Fig. 14c, the oxidation rates did not slow with cycle number, and in
tervals at a flow rate of 100 mLN·min−1. T = 1623 K was selected to fact the oxidation of cycles 2–9 was consistently more rapid than cycle
allow for the thermodynamic favorability of both cycle steps for the 1, as the majority of αn values were above the identity line.
given change in pO2 , and to achieve rapid reaction kinetics. The relative
temporal mass change Δm/m0 is shown in Fig. 13, and a dotted line 5. Solid characterization
indicates the theoretical Δm/m0 for full conversion. Including an initial
bakeout (cycle 0), the sample underwent 10 redox cycles and a cool- X-ray diffractometry (XRD, PANalytical X’Pert PRO Alpha-1 dif-
down in an oxidizing environment, shown in gray. The sample did not fractometer) was performed on reacted and unreacted samples of Fe2O3
fully re-oxidize during the bakeout cycle nor in any cycle after, con- and Fe3O4 at diffraction angles 2θ of 20–70° to verify sample compo-
sistent with thermodynamic equilibrium calculations (ΔG = 0) which sitions. Fig. 15 shows the intensity peaks for both species, with the
forecast a Fe3O4–Fe2O3 mixture for pO2 = 0.2 bar and T > 1100 K. unreacted (a) Fe3O4 and (b) Fe2O3 sample peaks shifted up for com-
While there was some variation in the oxidation extents per cycle due to parison with the corresponding (a) thermally reduced and (b) oxidized
slowed reaction rates as the sample neared thermodynamic equili- samples, respectively. The unreacted sample crystalline structures were
brium, there was no apparent systematic change in redox capacity first verified via comparison to star-quality XRD peaks from the PDF4+
during cycles 1–9. database (ICDD, 2016). Reacted samples were then compared to their
Fig. 14 depicts (a) the relative temporal sample mass changes for unreacted form to confirm complete thermal reduction and/or oxida-
cycles 0, 1, 5, and 9 superimposed for comparison and parity plots tion. Good agreement between sample peak locations and relative in-
showing α from cycles 3, 5, 7, and 9 as a function of α from cycle 1 for tensities was observed for both oxidation states with their unreacted
(b) thermal reduction and (c) oxidation, respectively. Both Fe2O3 counterparts, without apparent 2θ shifting or evidence of intermediate
thermal reduction and initial Fe3O4 oxidation proceeded rapidly, con- phases.
sistent with the kinetic analyses. During the thermal reduction stage, In-situ XRD (PANalytical Empyrean with Anton-Paar 1200N Hot
full conversion was achieved, supporting the assumption of complete Stage) was also performed to examine changes in the crystalline
thermal reduction in the solar reactor for the thermodynamic analysis. structures of oxidized Fe3O4 samples at elevated temperatures. Non-
For the re-oxidation stage, the final conversion varied significantly from reacting samples of Fe2O3 were heated to 1073 K in an atmospheric
cycle to cycle, with a maximum near α = 0.8. However, the sample did environment with diffraction peaks obtained in intervals of 200 K. The
approach α = 1 during the oxidizing cooldown, as shown in the gray resulting structures were compared to PDF4+ database entries for both
region (t > 18000 s) of Fig. 13, as the full oxidation of Fe3O4 to Fe2O3 -Fe2O3 and -Fe2O3 and are shown in Fig. 16. While there was some
became thermodynamically favorable at lower temperatures. According evidence of peak shifting, particularly at higher 2θ peaks, to lower

Fig. 13. Sample temperature (red) and mass


change relative to initial mass (solid black) for
redox cycling of Fe2O3 compared to the theore-
tical value (dotted black), driven by isothermal
O2 partial pressure swings (green), where the
gray region denotes cooldown in an oxidizing
environment. (For interpretation of the refer-
ences to colour in this figure legend, the reader is
referred to the web version of this article.)

624
H.E. Bush, P.G. Loutzenhiser Solar Energy 174 (2018) 617–627

Fig. 14. For initially-Fe2O3 samples, (a) superimposed reduction-oxidation cycle mass changes relative to initial sample mass; parity plots for (b) thermal reduction
and (c) re-oxidation comparing conversion to the first cycle.

Fig. 15. Intensity as a function of 2θ angle from X-ray diffractometry measurements for (a) an unreacted Fe3O4 sample and Fe3O4 produced via thermal reduction of
Fe2O3, (b) an unreacted Fe2O3 sample and Fe2O3 produced from oxidation of Fe3O4.

used to study particle size and morphology for the Fe2O3 and Fe3O4
samples (a) before and (b) after thermal reduction, shown in Fig. 17,
and oxidation, shown in Fig. 18, respectively. The majority of unreacted
Fe2O3 particles were much smaller than the 5 μm maximum specified
diameter, with smooth surfaces and near spherical shapes. Following
thermal reduction, the particle sizes grew significantly, resulting in less
spherical, slightly porous particles. The unreacted Fe3O4 particles had a
relatively wide range of particle sizes despite sieving, as well as more
angular initial surface shapes. Oxidation resulted in a porous, angular
particle sample in which the smaller particles were no longer present.
However, the apparent particle size growth was less significant relative
to the initially-Fe2O3 sample.

6. Conclusions

Thermodynamic and kinetic analyses were performed to evaluate


two-step, Fe2O3/Fe3O4 reduction-oxidation reactions coupled to an Air
Brayton cycle for thermochemical energy storage and electricity pro-
Fig. 16. Intensity as a function of 2θ angle from X-ray diffractometry mea- duction. Due to high Fe2O3 thermal reduction temperatures, increasing
surements of a non-reacting Fe2O3 sample between temperatures of 299 and
the solar concentration ratio to 4000 suns mitigated radiative losses and
1073 K.
provided a cycle efficiency of 46.0%, comparable to Co3O4/CoO at
equivalent conditions. Fe2O3 thermal reduction was described by an
angles, phase transition to -Fe2O3, indicated by loss of the first high- Avrami-Erofe’ev nucleation model of order n = 1.26 and a power-law
intensity peak between 30° < 2θ < 40°, was not observed. It was, O2 partial pressure dependency. Non-isothermal Fe3O4 oxidation ex-
therefore, concluded that the Fe3O4 samples from the kinetic analysis periments revealed two reaction rate peaks, suggesting a mechanism
were oxidized to -Fe2O3, consistent with the proposed oxidation transition consistent with known oxidation pathways. Modeling did not
pathway hypothesized in Fig. 10c–d. fully capture the transition, and further studies of early oxidation me-
Scanning electron microscopy (SEM, Zeiss Ultra 60 FE-SEM) was chanisms are required. Higher-temperature isothermal Fe3O4 analyses

625
H.E. Bush, P.G. Loutzenhiser Solar Energy 174 (2018) 617–627

Fig. 17. Scanning electron microscopy images for (a) an unreacted Fe2O3 sample and (b) Fe3O4 produced via thermal reduction of Fe2O3.

Fig. 18. Scanning electron microscopy images for (a) an unreacted Fe3O4 sample and (b) Fe2O3 produced via oxidation of Fe3O4.

revealed that significant oxidation within 20 s may be achieved. thermal radiative flux. Energy Fuels 27 (8), 4884–4890.
Pressure-swing Fe2O3/Fe3O4 cycling showed reduced conversion fol- Alonso, E., Pérez-Rábago, C., Licurgo, J., Fuentealba, E., Estrada, C.A., 2015. First ex-
perimental studies of solar redox reactions of copper oxides for thermochemical
lowing a bakeout thermal reduction due to thermodynamic limitations, energy storage. Sol. Energy 115, 297–305.
but the redox capacity did not systematically change over the following Alper, J.S., Gelb, R.I., 1990. Standard errors and confidence intervals in nonlinear re-
nine cycles. SEM revealed significant Fe2O3 particle size growth and an gression: comparison of Monte Carlo and parametric statistics. J. Phys. Chem. 94
(11), 4747–4751.
apparent decrease in surface area during thermal reduction. André, L., Abanades, S., Cassayre, L., 2017. High-temperature thermochemical energy
The combined thermodynamic-kinetic analysis of a solar thermo- storage based on redox reactions using Co-Fe and Mn-Fe mixed metal oxides. J. Solid
chemical cycle based on Fe2O3/Fe3O4 redox reactions coupled to an Air State Chem. 253 (Suppl. C), 6–14.
Babiniec, S.M., Coker, E.N., Miller, J.E., Ambrosini, A., 2015. Investigation of
Brayton cycle indicated superior maximum theoretical efficiencies for LaxSr1−xCoyM1−yO3−δ (M = Mn, Fe) perovskite materials as thermochemical energy
concentrating solar infrastructure capable of very high solar con- storage media. Sol. Energy 118, 451–459.
centration ratios. Kinetic analyses and cycling demonstrated rapid Babiniec, S.M., Coker, E.N., Miller, J.E., Ambrosini, A., 2016. Doped calcium manganites
for advanced high-temperature thermochemical energy storage. Int. J. Energy Res. 40
conversion at representative temperature and pressure conditions for
(2), 280–284.
both reactions. This work is an important prerequisite for guiding the Becker, K.D., von Wurmb, V., Litterst, F.J., 1993. Disorder and diffusional motion of
design and optimization of solar thermochemical reactors and non-solar cations in magnetite, Fe3−δO4, studied by Mössbauer spectroscopy at high tem-
re-oxidizers. These results establish an important upper bound on the peratures. J. Phys. Chem. Solids 54 (8), 923–935.
Block, T., Knoblauch, N., Schmücker, M., 2014. The cobalt-oxide/iron-oxide binary
performance of future implementations and inform the design and op- system for use as high temperature thermochemical energy storage material.
timization of thermochemical infrastructure. Thermochimica Acta 577 (Suppl. C), 25–32.
Block, T., Schmücker, M., 2016. Metal oxides for thermochemical energy storage: a
comparison of several metal oxide systems. Sol. Energy 126 (Suppl. C), 195–207.
Acknowledgements Bowrey, R.G., Jutsen, J., 1978. Energy storage using the reversible oxidation of barium
oxide. Sol. Energy 21 (6), 523–525.
This material is based upon work supported by the National Science Brendelberger, S., von Storch, H., Bulfin, B., Sattler, C., 2017. Vacuum pumping options
for application in solar thermochemical redox cycles – assessment of mechanical-, jet-
Foundation Graduate Research Fellowship under Grant No. DGE- and thermochemical pumping systems. Sol. Energy 141, 91–102.
1650044. Any opinions, findings, and conclusions or recommendations Carrillo, A.J., Sastre, D., Serrano, D.P., Pizarro, P., Coronado, J.M., 2016. Revisiting the
expressed in this material are those of the author(s) and do not ne- BaO2/BaO redox cycle for solar thermochemical energy storage. PCCP 18 (11),
8039–8048.
cessarily reflect the views of the National Science Foundation. The
Carrillo, A.J., Serrano, D.P., Pizarro, P., Coronado, J.M., 2014. Thermochemical heat
authors wish to thank Andrew Schrader for access to and discussion of storage based on the Mn2O3/Mn3O4 redox couple: influence of the initial particle size
his computational models for adaptation in this work. on the morphological evolution and cyclability. J. Mater. Chem. A 2 (45),
19435–19443.
Deutsch, M., Horvath, F., Knoll, C., Lager, D., Gierl-Mayer, C., Weinberger, P., Winter, F.,
References 2017. High-temperature energy storage: kinetic investigations of the CuO/Cu2O re-
action cycle. Energy Fuels 31 (3), 2324–2334.
Agrafiotis, C., Roeb, M., Schmuecker, M., Sattler, C., 2014. Exploitation of thermo- Feitknecht, W., Gallagher, K.J., 1970. Mechanisms for the oxidation of Fe3O4. Nature 228,
chemical cycles based on solid oxide redox systems for thermochemical storage of 548.
solar heat. Part 1: Testing of cobalt oxide-based powders. Sol. Energy 102, 189–211. Friedman, H.L., 1964. Kinetics of thermal degradation of char-forming plastics from
Alonso, E., Hutter, C., Romero, M., Steinfeld, A., Gonzalez-Aguilar, J., 2013. Kinetics of thermogravimetry. Application to a phenolic plastic. J. Polym. Sci. Part C: Polym.
Mn2O3–Mn3O4 and Mn3O4–MnO redox reactions performed under concentrated Symp. 6 (1), 183–195.

626
H.E. Bush, P.G. Loutzenhiser Solar Energy 174 (2018) 617–627

Gallagher, K.J., Feitknecht, W., Mannweiler, U., 1968. Mechanism of oxidation of mag- and modeling. Sol. Energy 150, 584–595.
netite to γ-Fe2O3. Nature 217, 1118. Schrader, A.J., Muroyama, A.P., Loutzenhiser, P.G., 2015. Solar electricity via an Air
Galvez, M., Loutzenhiser, P., Hischier, I., Steinfeld, A., 2008. CO2 splitting via two-step Brayton cycle with an integrated two-step thermochemical cycle for heat storage
solar thermochemical cycles with Zn/ZnO and FeO/Fe3O4 redox reactions: thermo- based on Co3O4/CoO redox reactions: thermodynamic analysis. Sol. Energy 118,
dynamic analysis. Energy Fuels 22 (5), 3544–3550. 485–495.
Gorman, B.T., Johnson, N.G., Miller, J.E., Stechel, E.B., 2015. Thermodynamic Shebanova, O.N., Lazor, P., 2003. Raman study of magnetite (Fe3O4): LASER-induced
Investigation of Concentrating Solar Power With Thermochemical Storage. (56840), thermal effects and oxidation. J. Raman Spectrosc. 34 (11), 845–852.
V001T005A026. Sibieude, F., Ducarroir, M., Tofighi, A., Ambriz, J., 1982. High temperature experiments
Halmann, M., Steinfeld, A., 2006. Fuel saving, carbon dioxide emission avoidance, and with a solar furnace: the decomposition of Fe3O4, Mn3O4, CdO. Int. J. Hydrogen
syngas production by tri-reforming of flue gases from coal- and gas-fired power Energy 7 (1), 79–88.
stations, and by the carbothermic reduction of iron oxide. Energy 31 (15), Sidhu, P.S., Gilkes, R.J., Posner, A.M., 1977. Mechanism of the low temperature oxidation
3171–3185. of synthetic magnetites. J. Inorg. Nucl. Chem. 39 (11), 1953–1958.
Haseli, P., Jafarian, M., Nathan, G.J., 2017. High temperature solar thermochemical SIGMA-ALDRICH, 2017. Safety Data Sheet: cobalt(II) oxide [1308-06-1]. Sigma-Aldrich,
process for production of stored energy and oxygen based on CuO/Cu2O redox re- Inc, Darmstadt, Germany.
actions. Sol. Energy 153, 1–10. Sahoo, S.K., Agarwal, K., Singh, A.K., Polke, B.G., Raha, K.C., 2010. Characterization of γ-
Hutchings, K.N., Wilson, M., Larsen, P.A., Cutler, R.A., 2006. Kinetic and thermodynamic and α-Fe2O3 nano powders synthesized by emulsion precipitation-calcination route
considerations for oxygen absorption/desorption using cobalt oxide. Solid State Ion. and rheological behaviour of α-Fe2O3. Int. J. Eng. Sci. Technol. 2 (8), 118–126.
177 (1), 45–51. Smith, K.S., Huyck, H.L., 1999. An overview of the abundance, relative mobility, bioa-
ICDD, 2016. PDF-4+ 2016. International Centre for Diffraction Data, Newtown Square, vailability, and human toxicity of metals. The Environ. Geochem. Mineral Deposits 6,
PA, USA. 29–70.
Imponenti, L., Albrecht, K.J., Wands, J.W., Sanders, M.D., Jackson, G.S., 2017. Stamatiou, A., Loutzenhiser, P.G., Steinfeld, A., 2010. Solar syngas production via H2O/
Thermochemical energy storage in strontium-doped calcium manganites for con- CO2-splitting thermochemical cycles with Zn/ZnO and FeO/Fe3O4 redox reactions.
centrating solar power applications. Sol. Energy 151, 1–13. Chem. Mater. 22 (3), 851–859.
Kissinger, H.E., 1957. Reaction kinetics in differential thermal analysis. Anal. Chem. 29 Steinfeld, A., Fletcher, E.A., 1991. Theoretical and experimental investigation of the
(11), 1702–1706. carbothermic reduction of Fe2O3 using solar-energy. Energy 16 (7), 1011–1019.
Linstrom, P.J., Mallard, W.G., 2001. The NIST Chemistry WebBook: a chemical data re- Steinfeld, A., Frei, A., Kuhn, P., 1995. Thermoanalysis of the combined Fe3O4-reduction
source on the internet. J. Chem. Eng. Data 46 (5), 1059–1063. and CH4-reforming processes. Metall. Mater. Trans. B 26 (3), 509–515.
Loutzenhiser, P.G., Gálvez, M.E., Hischier, I., Stamatiou, A., Frei, A., Steinfeld, A., 2009. Steinfeld, A., Kuhn, P., Karni, J., 1993. High-temperature solar thermochemistry: pro-
CO2 splitting via two-step solar thermochemical cycles with Zn/ZnO and FeO/Fe3O4 duction of iron and synthesis gas by Fe3O4-reduction with methane. Energy 18 (3),
redox reactions II: kinetic analysis. Energy Fuels 23 (5), 2832–2839. 239–249.
Monazam, E.R., Breault, R.W., Siriwardane, R., 2014. Kinetics of magnetite (Fe3O4) Steinfeld, A., Kuhn, P., Reller, A., Palumbo, R., Murray, J., Tamaura, Y., 1998. Solar-
oxidation to hematite (Fe2O3) in air for chemical looping combustion. Ind. Eng. processed metals as clean energy carriers and water-splitters. Int. J. Hydrogen Energy
Chem. Res. 53 (34), 13320–13328. 23 (9), 767–774.
Muroyama, A.P., Schrader, A.J., Loutzenhiser, P.G., 2015. Solar electricity via an Air Steinfeld, A., Palumbo, R., 2001. Solar thermochemical process technology. Encyclopedia
Brayton cycle with an integrated two-step thermochemical cycle for heat storage Phys. Sci. Technol. 15 (1), 237–256.
based on Co3O4/CoO redox reactions II: Kinetic analyses. Sol. Energy 122, 409–418. Tamaura, T., Tsuji, M., Ehrensberger, K., Steinfeld, A., 1997. The solar-driven coal/Fe3O4
Nakamura, T., 1977. Hydrogen production from water utilizing solar heat at high tem- redox system. J. Phys. IV France 07 (C1) C1-685-C681-686.
peratures. Sol. Energy 19 (5), 467–475. Tiernan, M.J., Barnes, P.A., Parkes, G.M.B., 2001. Reduction of iron oxide catalysts: the
Neises, M., Tescari, S., de Oliveira, L., Roeb, M., Sattler, C., Wong, B., 2012. Solar-heated investigation of kinetic parameters using rate perturbation and linear heating ther-
rotary kiln for thermochemical energy storage. Sol. Energy 86 (10), 3040–3048. moanalytical techniques. J. Phys. Chem. B 105 (1), 220–228.
Pagkoura, C., Karagiannakis, G., Zygogianni, A., Lorentzou, S., Kostoglou, M., Vyazovkin, S., Burnham, A.K., Criado, J.M., Pérez-Maqueda, L.A., Popescu, C.,
Konstandopoulos, A.G., Rattenburry, M., Woodhead, J.W., 2014. Cobalt oxide based Sbirrazzuoli, N., 2011. ICTAC Kinetics Committee recommendations for performing
structured bodies as redox thermochemical heat storage medium for future CSP kinetic computations on thermal analysis data. Thermochim. Acta 520 (1), 1–19.
plants. Sol. Energy 108, 146–163. Wong, B., 2011. Thermochemical heat storage for concentrated solar power, final report
Papanastassiou, D., Bitsianes, G., 1973. Mechanisms and kinetics underlying the oxida- for the US Department of Energy. San Diego, CA, USA.
tion of magnetite in the induration of iron ore pellets. Metall. Trans. 4 (2), 487–496. Yu, J., Han, Y., Li, Y., Gao, P., Li, W., 2017. Mechanism and kinetics of the reduction of
Parkinson, G.S., 2016. Iron oxide surfaces. Surf. Sci. Rep. 71 (1), 272–365. hematite to magnetite with CO–CO2 in a micro-fluidized bed. Minerals 7 (11), 209.
Schrader, A.J., De Dominicis, G., Schieber, G.L., Loutzenhiser, P.G., 2017. Solar electricity Zhang, Z., Andre, L., Abanades, S., 2016. Experimental assessment of oxygen exchange
via an Air Brayton cycle with an integrated two-step thermochemical cycle for heat capacity and thermochemical redox cycle behavior of Ba and Sr series perovskites for
storage based on Co3O4/CoO redox reactions III: Solar thermochemical reactor design solar energy storage. Sol. Energy 134, 494–502.

627

You might also like