You are on page 1of 13

A Novel CeO2–xSnO2/Ce2Sn2O7 Pyrochlore Cycle for Enhanced

Solar Thermochemical Water Splitting


Chongyan Ruan and Yuan Tan
State Key Laboratory of Catalysis, Dalian Institute of Chemical Physics, Chinese Academy of Sciences,
Dalian 116023, China
University of Chinese Academy of Sciences, Beijing 100049, China
Lin Li, Junhu Wang, Xiaoyan Liu, and Xiaodong Wang
State Key Laboratory of Catalysis, Dalian Institute of Chemical Physics, Chinese Academy of Sciences,
Dalian 116023, China

DOI 10.1002/aic.15701
Published online March 17, 2017 in Wiley Online Library (wileyonlinelibrary.com)

A novel CeO2–xSnO2/Ce2Sn2O7 pyrochlore stoichiometric redox cycle with superior H2 production capacities is identi-
fied and corroborated for two-step solar thermochemical water splitting (STWS). During the first thermal reduction step
(14008C), a reaction between CeO2 and SnO2 occurred for all the CeO2–xSnO2 (x 5 0.05–0.20) solid compounds, form-
ing thermodynamically stable Ce2Sn2O7 pyrochlore rather than metastable CeO2-d. Consequently, substantially higher
reduction extents were achieved owing to the reduction of CeIV to CeIII. Moreover, in the subsequent reoxidation with
H2O (8008C), H2 production capacities increased by a factor of 3.8 as compared to the current benchmark material
ceria when x 5 0.15, with the regeneration of CeO2 and SnO2 and the concomitant reoxidation of CeIII to CeIV. The
H2O-splitting performance for CeO2–0.15SnO2 was reproducible over seven consecutive redox cycles, indicating the
C 2017 American Institute of Chemical Engineers AIChE J, 63: 3450–3462, 2017
material was also robust. V
Keywords: solar fuel, pyrochlore, stoichiometric, water splitting, H2 production

Introduction High-temperature reduction


Solar resource, with its unmatched magnitude and availabil- 1
MOox ! MOred 1 O2 ðgÞ (1)
ity, is generally considered as a permanent solution to ensuring 2
energy security and combating climate change.1–3 Currently, Low-temperature oxidation with H2O
several plausible pathways to convert sunlight to fuel inter-
mediates, including photochemical,4,5 photoelectrochemi- MOred 1H2 OðgÞ ! MOox 1H2 ðgÞ (2)
cal,6,7 thermochemical,8–10 and their combinations11,12 are
where MOox and MOred represent the oxidized and the corre-
under consideration. In particular, solar thermochemical water
sponding reduced forms of the metal oxide.
splitting (STWS), which utilizes the entire solar spectrum and
During the first high-temperature reduction step, MOox is
inherently operates at high temperatures, offers the potential
typically reduced at low oxygen partial pressure driven by
to convert abundant yet intermittent solar energy to renewable
concentrated solar energy. During the following oxidation
hydrogen with high solar to hydrogen production rates and
efficiencies.8–10,13–16 step, H2O is introduced to reoxidize the reduced metal oxide
Two-step thermochemical cycles, which eliminate the need back to the original oxidized state, producing hydrogen and
for in situ gaseous separation of O2 and H2,17–21 and lower the thus establishing a cyclic process.
required temperature compared to direct water thermoly- Over the past decades, numerous efforts have been made to
sis,13,22 are most commonly operated in STWS. A typical two- identify viable materials for efficient two-step
step STWS cycle involving metal oxide as reactive intermedi- STWS.14–16,23–25 However, substantial challenges still remain
ate can be exemplified as follows: in discovering feasible metal-oxide based redox cycles with
moderate operating temperatures and large gravimetric fuel
production capacities for cost-effective solar hydrogen pro-
duction. In early studies, systems based on stoichiometric
Additional Supporting Information may be found in the online version of this
article.
chemistries (e.g., Fe3O4 ! 3FeO 1 1=2O226,27; ZnO !
Zn 1 1=2O228,29; SnO2 ! SnO 1 1=2O230,31) were extensive-
Correspondence concerning this article should be addressed to X. Wang at ly studied on account of the great oxygen exchange capacities
xdwang@dicp.ac.cn.
and thus fuel productions per mass of oxide. Cycle feasibility
C 2017 American Institute of Chemical Engineers
V with pure Fe3O4 has been experimentally demonstrated, but

3450 August 2017 Vol. 63, No. 8 AIChE Journal


the kinetics are slow and magnetite must be combined with an the oxygen-vacancy formation energy and in consequence
oxygen conducting refractory material such as yttrium- enhances the reduction extents. Among them Hf-doped ceria
stabilized zirconia (YSZ) or ZrO2 to suppress problems associ- exhibits the best redox performance in term of the fuel produc-
ated with melting and sintering, which limit their overall tion for CO and H2, with 35% more CO and 41% more H2
solar-to-fuel efficiency. Conversely, for the ZnO/Zn and evolving for Ce0.9Hf0.1O2 than undoped ceria.36,42 Moreover,
SnO2/SnO processes, the volatile products require rapid in a recent study, rhodium-doped ceria was employed in solar
quenching to avoid recombination and to achieve products thermochemical cycles with stable solar H2 and even hydro-
separation, yielding great challenges for innovative reactor carbon fuels productivity by incorporating a catalytic pro-
design even though the global solar-to-fuel efficiency is cess.40 Despite the extensive effort, the above doping metals
promising.24 based on nonstoichiometric chemistries have achieved limited
Recently, two-step STWS cycles based on nonstoichiomet- success in increasing H2 production capacities due to the par-
ric chemistries has attracted considerable interest for the sim- tial reduction of CeIV to CeIII. In previous studies, Sn doped
plicity of implementation. Three classes of nonstoichiometric ceria solid solutions exhibited numerous applications as oxy-
oxides have been proposed as candidate materials for commer- gen gas sensors and oxygen ion conductors by virtue of their
cial two-step STWS: doped-hercynite (CoxFe1–xAl2O4), perov- amazing oxygen exchange properties, which relied on the
skite (SrxLa1–xMnyAl1–yO3), and ceria (CeO2). Muhich et al.19 redox cycle between CeIV and CeIII within the Ce2Sn2O7-
validated the doped-hercynite cycle governed by an O- Ce2Sn2O8 complex oxides known as pyrochlores.46 The pyro-
vacancy mechanism using DFT calculations in conjunction chlore oxides associated with the general formula A2B2O7 can
with experimental measurements. Near-isothermal H2O split- be described as cubic fluorite consisting of ordered cations and
ting was achieved by cycling CoxFe1–xAl2O4 materials vacancies with 1/8 of the anions missing,47 which have been
between 1500 and 13508C. Slow reaction kinetics is currently the subject of great interest motivated by the wide variety of
the most significant limitation for the doped-hercynite active the chemical substitutions at the A, B, O sites, significant
materials. Recent studies on perovskites (ABO3) highlight this oxide ion conductivity and radiation tolerance under extreme
class of oxides as impressive candidates for solar-to-fuel con- conditions.48–54 Therefore, it would be highly anticipated that
version. McDaniel et al.20 evaluated the nonstoichiometric the CeO2–xSnO2 solid compounds can potentially enhance the
SrxLa1–xMnyAl1–yO3–d perovskites and demonstrated encour- reduction extents and thus improve two-step STWS perfor-
aging fuel productivity at a reduction temperature of 13508C. mance. However, the lack of studies related to CeO2–xSnO2
Based on the highly amenability to doping, perovskites with solid compounds in two-step STWS, in particular, is surprising
higher H2 production capacity and more rapid reaction kinetics given the superior oxygen exchange capacities.
still remain to be identified for practical solar fuel applica- In light of this evidence, we synthesized a series of ceria-tin
tions. Undoped ceria has emerged as a highly attractive candi- oxide solid compounds (CeO2–xSnO2) for assessment of their
date with inherent fast reaction kinetics and superior two-step STWS capabilities as compared to benchmark mate-
crystallographic stability, making it the current benchmark rial ceria. The productivity and overall reaction kinetics of the
material for comparing the performance of new redox cycles prepared materials for two-step STWS cycled between 1400
for two-step STWS.8,21,32,33 Chueh et al.8,33 proposed a non- and 8008C were systematically investigated in identical exper-
stoichiometric ceria cycle via an O-vacancy mechanism as imental conditions. In parallel, crystalline phase evolution,
illustrated below: element-specific chemical state, microstructural morphology
Higher temperature change, and elemental composition distributions of the sam-
d ples conditioned under various redox treatment (initial sam-
CeO2 ! CeO22d 1 O2 (3) ples, samples subjected to thermal reduction, postcycling
2
samples), were characterized and critically analyzed via pow-
Lower temperature der X-ray diffraction (XRD), Raman spectroscopy, X-ray
CeO22d 1dH2 O ! CeO2 1dH2 (4) absorption near edge structure (XANES) spectroscopy,
M€ossbauer spectroscopy, and electron microscopy (SEM,
The viability of the nonstoichiometric ceria cycle was success- HRTEM, STEM-EDX). Finally, insights into reaction mecha-
fully demonstrated with stable solar fuels production in a nism governing the two-step redox cycle were deduced based
cavity-based solar reactor under realistic solar concentrating on the experimental findings and characterization evidence.
conditions.8,21,34 Nevertheless, the reduction to nonstoichio- Essentially, a new CeO2–xSnO2/Ce2Sn2O7 pyrochlore stoi-
metric CeO22d shows restricted thermochemical fuel yield chiometric redox cycle taking advantage of reversible change
despite the fast kinetics and outstanding thermal stability. of the oxidation state of Ce, that is, CeIV $ CeIII was identi-
Rather high temperatures, near 20008C, were necessary to fied to facilitate two-step STWS.
decompose CeIVO2 to CeIII 2 O3 for efficient fuel production,
35

which impeded the practicality of undoped ceria in STWS Experimental Section


applications.
Doping ceria with divalent (Ca, Mg, Sr, Cu),36,37 trivalent Preparation of redox materials
(Y, La, Sc, Sm, Gd, Pr, Fe, Ni, Rh),32,38–40 and tetravalent (Zr, All the chemicals used in the coprecipitation synthesis of
Hf)41–45 cations to form CeO2-based mixed oxides or solid sol- CeO2–xSnO2 (x 5 0.05, 0.10, 0.15, and 0.20, i.e., molar ratio
utions has proven to be an effective way to alter its thermody- of Ce:Sn 5 0.95:0.05, 0.90:0.10, 0.85:0.15, and 0.80:0.20) sol-
namic and kinetic properties. The divalent and trivalent id compounds were of analytical grade without further treat-
dopants introduce stable intrinsic oxygen vacancies within the ment. Typically, stoichiometric amounts of Ce(NO3)36H2O
ceria lattice and thus result in higher oxygen diffusion rates and SnCl22H2O were mixed and dissolved in deionized water
when codoped with tetravalent dopants. The tetravalent dop- under continuous stirring. To dissolve SnCl22H2O precursor,
ants induce lattice strain in bulk ceria, which in turn decreases dilute hydrochloric acid was added slowly to the magnetically

AIChE Journal August 2017 Vol. 63, No. 8 Published on behalf of the AIChE DOI 10.1002/aic 3451
Table 1. Nominal and Measured Material Compositions of Reactivity tests for thermochemical cycling
CeO2–xSnO2 Solid Compounds with Associated Sample
The experimental setup for two-step water splitting is
Names
shown schematically in Figure 1. Powder Sample (0.5 g) was
Nominal Ce:Sn Measured Ce:Sn loosely packed in a stagnation flow reactor as reported else-
Sample Name Molar Ratio Molar Ratio where,9,18,55 allowing gases to access all the redox material.
CeO2-0.05SnO2 0.95:0.05 0.96:0.04 The main features of the reactor are briefly summarized here.
CeO2-0.10SnO2 0.90:0.10 0.92:0.08 A stainless gas-handing manifold, two concentric Al2O3 tubes
CeO2-0.15SnO2 0.85:0.15 0.85:0.15 make up the reactor. Inlet gases flow along the open ended
CeO2-0.20SnO2 0.80:0.20 0.79:0.21
inner tube (7 mm i.d.) downward toward the round-bottom,
close ended outer tube (14 mm i.d.) to mix well with the mate-
stirred solution until pH 5 1 to obtain a transparent solution. rial tested, and then turn by 1808 to exit via an annular region
Subsequently, ammonium hydroxide was added dropwise to between the two concentric tubes. The concentric alumina
the transparent solution until pH 5 9, under vigorously stirring tubes reactor was placed horizontally in the infrared image
for 2 h at room temperature. The resultant slurry was aged for furnace (ULVAC-RIKO, VTH-E44) capable of rapid heating
another 1 h, filtered, washed thoroughly with deionized water, and cooling. Irradiation from the infrared image lamps was
and then dried at 1208C overnight. Finally, the dried solid regulated by a temperature program controller (ULVAC-
products were ground into a fine powder with an agate mortar RIKO, TPC-5000), and temperature was monitored using an
and pestle followed by thermal treatment in the flowing air R-type thermocouple (63.88C) enclosed in an alumina sheath
according to the program: heating from room temperature to in direct contact with the sample.
10008C at 58C min21, and finally heating to 14008C at 28C Gas flows were regulated with digital mass flow controllers.
min21 and holding for 4 h. This resulted in sintered, light yel- Prior to entering the reactor, traces of O2 in purge gas Ar
low dense powders. For comparison, pure CeO2 sample was (99.9996% purity) was removed by passing through the deoxi-
prepared following the same procedure as described above dizer. During a typical thermochemical cycle, thermal reduc-
without adding SnCl22H2O precursor. It must be mentioned tion was first accomplished by increasing the sample
that the sintering process at the high temperature relevant to temperature from 8008C to 14008C at 1008C min21 in an Ar
thermal cycling is crucially important to stabilize the structure flow (500 mL min21), and holding at 14008C for 10 min. Fol-
properties of the samples as prepared. Fluctuations of the sam- lowing thermal reduction, the temperature was ramped down
ples’ performance during two-step STWS can thus be to 8008C at 2008C min21 in the flowing Ar. Subsequently,
minimized. H2O splitting was initiated by delivering Ar gas saturated with
A summary of desired molar ratio of Ce:Sn, actual molar purified water (500 mL min21, c(H2O) 5 4.2 mol %) to the
ratio of Ce:Sn as determined by inductively coupled plasma- thermally reduced sample via a water bubbler inside a water
atomic emission spectrometry (ICP-AES) analysis, and the bath at 308C for 8 min. The passage for the steam to the reac-
associated sample identifications are presented in Table 1. tor was heated by a heater band to prevent condensation. Both
Inductively coupled plasma2mass spectrometry (ICP-MS) the high-temperature reduction and low-temperature H2O
analysis detected no residual chlorine from the as synthesized splitting steps were carried out at atmospheric pressure.
CeO2–0.15SnO2 sample sintered at 14008C for 4 h (Supporting Product gases (O2, H2) exiting the flow reactor were con-
Information Table S1). stantly monitored and recorded using a quadrupole mass

Figure 1. Experimental setup for two-step thermochemical water splitting.


[Color figure can be viewed at wileyonlinelibrary.com]

3452 DOI 10.1002/aic Published on behalf of the AIChE August 2017 Vol. 63, No. 8 AIChE Journal
spectrometer (MS; InProcess Instruments GAM200) after con- least-squares method using the MossWinn 4.0 program. The
119
densing out of excess steam by a cryogenic trap. Calibration Sn isomer shifts were given relative to the center of the
of the mass spectrometry was performed daily prior to each CaSnO3 spectrum recorded at room temperature.
experiment with six known concentrations of O2 and H2 Surface microstructures were analyzed via scanning elec-
(1000–5000 ppm O2, 1000–5000 ppm H2, balance Ar) to an tron microscopy (JSM-7800F Field Emission SEM) operated
accuracy of 0.001 mol %. at 4 kV. Samples were placed on conductive tapes and then
The transient O2 and H2 evolution rates were calculated as sputtered with carbon. The morphology and crystallinity of the
xi Vtotal materials were also characterized using a JEOL JEM-2000
Vi 5 (5) EM field emission high-resolution transmission electron
mred
microscope (HRTEM) operating at 200 kV. Samples after var-
where Vi denotes the volumetric rates of O2 or H2 produced ious processing steps were crushed into fine powders, followed
per unit of mass of CeO2–xSnO2 sample; xi denotes the mole by ultrasonic dispersion in ethanol for 15 min. The resulting
fraction of O2 or H2 in the effluent gas, monitored by the mass suspension was drop-casted on copper TEM grids coated with
spectrometer; Vtotal is the total volumetric flow rate regulated carbon and allowed to dry for analysis. Elemental composition
by the digital mass flow controller; and mred is the mass of and distribution were determined via energy dispersive X-ray
CeO2–xSnO2 redox material. spectroscopy (EDX).
For quantitative determination of effluent gas, O2 and H2 Nitrogen sorption isotherm at 77 K was performed on a
yields per cycle were obtained from numerical integration of Micromeritics ASAP 2460 apparatus. The samples were out-
gas evolution rate-time curves. gassed at 3008C and 1-m Torr for 4 h prior to the analysis.
Brunauer-Emmett-Teller (BET) were evaluated from the
Characterization techniques adsorption isotherm within the pressure range 0.05 < P/P0<
At the end of each processing step, the redox material was 0.3.
cooled down under Ar atmosphere. Afterward, the powder
was transferred into an air-closed container under N2-atmo- Results and Discussion
sphere for ex situ structure characterization. The crystal phase Thermal chemical H2O splitting
transformation of CeO2–xSnO2 (x 5 0.05–0.20) redox samples
(samples as-synthesized, samples subjected to thermal reduc- For thermochemical cycling applications, pure CeO2 acts as
tion, samples after H2O splitting reaction), were identified by a good baseline for comparison of new materials to evaluate
X-ray powder diffraction (XRD) using a PANalytical diffrac- the degree of enhancement in solar fuels productivity that can
tometer (Cu Ka, 40 kV, 40 mA). A continuous mode was used be achieved. We therefore compared the transient O2 and H2
for collecting data in the 2h scan range of 15–808 with an evolution rates during a complete H2O-splitting redox cycle
angular step size of 0.028 and a counting time of 30 s per step. obtained with CeO2–0.15SnO2 and pure CeO2 under identical
Raman spectroscopy was conducted using a LabRam experimental conditions. The results are presented in Figures
HR800 confocal Raman microprobe (HORIBA Jobin Yvon, 2a, b. During the thermal reduction step (Figure 2a), O2 evolu-
France) equipped with a 532-nm wavelength He-Ne laser. tion starts at above 9008C for both CeO2–0.15SnO2 and pure
Phonon modes in the powders were investigated to fingerprint CeO2. Intriguingly, O2 evolution from CeO2–0.15SnO2
the chemical structure of CeO2–xSnO2 (x 5 0.05–0.20) sam- increases more rapidly initially in comparison with pure CeO2,
ples after various processing steps. attaining a peak rate of 2.83 mL min21 g21 at 11008C and
X-ray absorption near edge structure (XANES) spectra of 0.30 mL min21 g21 at 14008C for CeO2–0.15SnO2 and pure
the samples (diluted with approximated 90% by mass of BN CeO2, respectively. Notably, fully 4.73 more O2 is released
and pressed into self-supporting wafers) at the Ce LIII-edge for CeO2–0.15SnO2 (4.87 mL g21) than pure CeO2 (1.03 mL
were measured at beamline BL14W1 of Shanghai Synchrotron g21). During the following oxidation with H2O at 8008C (Fig-
Radiation Facility (SSRF, Shanghai, China). The electron stor- ure 2b), H2 is immediately produced on introduction of the
age ring was operated at 3.5 GeV with the electron current of steam in both cases. Pure CeO2 exhibits an initial rapid
140 2 210 mA. Spectra were collected in transmission mode increase to a peak rate of 5.08 mL min21 g21 followed by
at room temperature using a solid-state detector. The output exponential decline. Whereas for CeO2–0.15SnO2, the peak
beam was monochromatized by a Si(111) double-crystal H2 evolution rate drops to 2.96 mL min21 g21 followed by a
monochromator. CeO2 and Ce(NO3)36H2O wafers were used comparably slower decrease rate. Conversely, compared to
as standard references with known oxidation states of CeIV pure CeO2, CeO2–0.15SnO2 did not achieve its thermodynam-
and CeIII, respectively. The XANES data were recorded from ic equilibrium, as indicates by long tail of H2 evolution after
5520 to 6120 eV and analyzed using a standard curve-fitting reoxidation at 8008C for 8 min. And it was due to the compa-
procedure with ATHENA software program. rably short reoxidation period (see Supporting Information
119
Sn Mo€ssbauer spectroscopic measurements were carried Figure S1 for further information). More importantly, the
out in transmission mode at room temperature using a constant enhanced reduction extent for CeO2–0.15SnO2 leads to a
acceleration spectrometer (Topologic Systems. Inc.) with a higher H2 yield during the subsequent oxidation with H2O, for
single-line c-ray source of Ca119mSnO3. The absorbers con- which H2 production capacity (CeO2–0.15SnO2, 7.18 mL g21;
taining approximately 10 mg cm22 of 119Sn were prepared pure CeO2, 1.87 mL g21) increases by a factor of 3.8 despite
under argon atmosphere inside the glove box, and the sample the initial H2 evolution rate being lower on introduction of the
holder was then sealed to avoid contact with air. Velocity steam.
scale calibration was performed using the magnetic sextet of a To get further insight into the influence of SnO2 content on
high-purity a-Fe foil absorber as a standard, and 57Co (Rh the thermal reduction capabilities and the subsequent H2O
matrix) was used as the source. Experimental spectra were fit- splitting kinetics, CeO2–xSnO2 (x 5 0.05–0.20) together with
ted to appropriate combinations of Lorentzien profiles by pure CeO2 were assessed systematically. The thermochemical

AIChE Journal August 2017 Vol. 63, No. 8 Published on behalf of the AIChE DOI 10.1002/aic 3453
splitting performances are also sufficiently favored owing to
the enhanced reduction extents.
To assess the stability of the water-splitting performance,
CeO2–0.15SnO2 was subjected to seven consecutive redox
cycles at 14008C reduction/8008C reoxidation. As seen in Fig-
ure 3b, the O2 and H2 productivity are roughly constant with-
out obvious deactivation from cycle 1 to cycle 7. The average
yields of O2 and H2 for CeO2–0.15SnO2 are 4.64 and 7.05 mL
g21, respectively, with very slightly decreases compared with
the values reported in Supporting Information Table S2, indi-
cating rather stable fuel productivity. ICP-AES analysis shows
the molar ratio of Ce:Sn for CeO2–0.15SnO2 sample after
cycling is consistent with that of initial sample sintered at
14008C for 4 h (Supporting Information Table S3).
To probe into the underlying reason for the enhanced reduc-
tion extents with sufficiently favored H2O splitting performan-
ces of ceria-tin oxide solid compounds when compared with

Figure 2. Comparison of transient (a) O2 evolution rates


during thermal reduction at 14008C in Ar, (b) H2
evolution rates during the following reoxidation
at 8008C in H2O/Ar for CeO2–0.15SnO2 and pure
CeO2.
Experimental conditions were as follows: Ar sweep gas
flow rate 5 1000 mL min21 g21 during thermal reduc-
tion; cH2O 5 4.2 mol %, total gas flow rate 5 1000 mL
min21 g21 during the following reoxidation. Inset shows
the integrated O2 and H2 productivity, respectively.
[Color figure can be viewed at wileyonlinelibrary.com]

behaviors are depicted in Figure 3a (see Supporting Informa-


tion Figure S2 and Table S2 for details). We note that for all
the CeO2–xSnO2 (x 5 0.05–0.20) solid compounds, total
amounts of oxygen evolved are improved significantly when
compared with pure CeO2, and increase linearly with increas-
ing SnO2 content. Given that all the samples were evaluated
under the same conditions, we thus presume the enhanced O2
yields mainly stem from more favorable reduction thermody-
namics. H2 yields, in excellent correspondence with the
expectations, also increase linearly with the improved reduc-
tion extent for x 5 0.05–0.15. It is noteworthy that H2 yields Figure 3. (a) O2 and H2 yield of CeO2–xSnO2 (x 5 0.05–
increase by a factor of 1.3 and even 3.8 for x 5 0.05 and 0.20) solid compounds as compared to pure
x 5 0.15, respectively (Supporting Information Figure S2 and CeO2. (b) O2 and H2 yield of CeO2–0.15SnO2
Table S2). However, for higher SnO2 content of x 5 0.20, H2 over seven consecutive H2O splitting cycles.
yield decreases remarkably albeit the highest O2 yield is The same experimental conditions were employed and
observed, which may result from the severe sintering as will illustrated as follows: Ar sweep gas at a flow rate of
be shown in the subsequent discussion. In summary, the reduc- 1000 mL21 min21 g21 during the thermal reduction at
14008C, and H2O/Ar at cH2O 5 4.2 mol %, for which
tion extents for CeO2–xSnO2 (x 5 0.05–0.20) are substantially total gas flow rate 5 1000 mL21 min21 g21 during the
improved in comparison with pure CeO2, and increase linearly reoxidation at 8008C. [Color figure can be viewed at
with increasing SnO2 content. Moreover, the subsequent H2O wileyonlinelibrary.com]

3454 DOI 10.1002/aic Published on behalf of the AIChE August 2017 Vol. 63, No. 8 AIChE Journal
pure CeO2, crystalline phase, element-specific chemical state,
microstructural morphology, and elemental composition distribu-
tions of the samples (initial samples, samples subjected to thermal
reduction, samples after the subsequent H2O splitting cycle) of
various processing steps, were characterized via XRD together
with Raman spectroscopy, XANES spectroscopy, M€ossbauer
spectroscopy, as well as electron microscopy (SEM, HRTEM,
STEM-EDX).
Sample characterization
Figure 4a displays the X-ray powder diffraction patterns of
initial CeO2–xSnO2 (x 5 0.05–0.20) solid compounds thermal-
ly calcined at 14008C for 4 h. Pure CeO2 is also shown as a
reference. For all the as synthesized samples, reflection peaks
indexable to the cubic fluorite CeO2 (JCPDS 00-043-1002)
and rutile type tetragonal SnO2 (JCPDS 01-077-0447) are
observed, indicating the samples prior to use are in the form of
mixed metal oxides. Additionally, the intensity of SnO2 dif-
fraction peaks, as expected, monotonically increases with
increasing SnO2 amount. The X-ray powder diffraction pat-
terns of CeO2–xSnO2 (x 5 0.05–0.20) solid compounds after
thermal reduction at 14008C in flowing Ar for 10 min are
shown in Figure 4b. Interestingly, for all the samples thermally
reduced, reflection peaks assignable to the SnO2 completely
disappear (with traces of SnO2 only detectable for x 5 0.20)
along with the emergence of a new superstructure reflection
attributable to the Fd3m (no. 227) symmetry of a typical
Ce2Sn2O7 pyrochlore phase. The newly emerged reflection
peaks (highlight by the gray bars in Figures 4b, c) can be
indexed as (222), (400), (440), and (622) planes of Ce2Sn2O7
pyrochlore structure.56 Furthermore, the increase of the corre-
sponding Ce2Sn2O7 pyrochlore reflections is coupled with the
decrease of the cubic fluorite CeO2 phase. This points out that
a reaction between CeO2 and SnO2 has occurred, resulting in
the reduction or even disappearance of the reflection peaks of
CeO2 and SnO2. No extra peaks originating from the other
phases are observed for pure CeO2 sample thermally reduced
(Figure 4b). The above XRD observations confirm that during
thermal reduction in flowing Ar, a stoichiometric reaction
occurs between CeO2 and SnO2, forming thermodynamically
favorable Ce2Sn2O7 pyrochlore. These observations are also in
agreement with the enhanced O2 production capacities, which
increase linearly with increasing SnO2 content as shown in
Figure 3a. Figure 4. X-ray powder diffraction patterns of CeO2–xSnO2
Notably, after the subsequent H2O splitting step with the solid compounds: (a) as synthesized samples
thermally reduced CeO2–xSnO2 (x 5 0.05–0.20) solid com- calcined at 14008C, (b) samples thermally
pounds, the intensities of the Ce2Sn2O7 pyrochlore reflection reduced in flowing Ar, (c) samples after the sub-
peaks decrease remarkably, with only traces of Ce2Sn2O7 sequent H2O splitting step.
pyrochlore phase present when x 5 0.05–0.15 (Figure 4c). Crystalline phases: (•) CeO2; (*) SnO2; (䉬) Ce2Sn2O7.
This reveals that the reduced Ce2Sn2O7 pyrochlore is almost The gray bars highlight reflection peaks assignable to
completely oxidized. From these results, it is concluded that Ce2Sn2O7 pyrochlore. Pure CeO2 after various process-
during reoxidation with H2O, Ce2Sn2O7 pyrochlore is oxidized ing steps are also shown as references. [Color figure can
be viewed at wileyonlinelibrary.com]
back to CeO2 and SnO2. Additionally, no diffraction peak
characteristic of SnO2 is resolved, indicating a high dispersion
of SnO2 in ceria-tin oxide solid compounds after the reoxida-
tion step. On the contrary, in the case of CeO2–xSnO2 (x 5 0.05–0.20) solid compounds. Raman spectra of samples
(x 5 0.20), Ce2Sn2O7 pyrochlore is still visible, thus implying after various processing steps are compared in Figure 5a (see
a fraction of the reduced Ce2Sn2O7 pyrochlore has not been Supporting Information Figures S3 and S4 for more details).
oxidized back by H2O, resulting in the remarkable decrease in The spectra are plotted on the same relative intensity scale.
H2 yield presented in Figure 3a. All the as synthesized samples calcined at 14008C exhibit
As an effective tool to analyze the M-O bonds as well as lat- spectral features indicative of ideal crystalline CeO2, with an
tice defects, Raman spectroscopy was employed to further intense band centered at ca. 465 cm21 (Figure 5a) related to
characterize the crystalline phase evolution of CeO2–xSnO2 the F2g Raman active mode of the fluorite-type structure

AIChE Journal August 2017 Vol. 63, No. 8 Published on behalf of the AIChE DOI 10.1002/aic 3455
for all the samples thermally reduced as compared to the as
synthesized samples, for which DF2g Raman shift is almost
linear with SnO2 molar ratio, Figure 5b. These phenomena are
qualitatively consistent with other Ce12xRExO22y systems,
where RE 5 La, Pr, Nd, Eu, Gd, and Sm.58–60,62–64 For exam-
ple, Banerji et al.59 and Giannici et al.63 reported that as the
M2O3 (M 5 Gd, Sm, Yb) concentration in CeO2 increases, the
Raman spectrum changes considerably with increases in width
and a decrease of the intense mode of CeO2 at 465cm21. In a
recent study, Li et al.65 also demonstrated that the insertion of
Sn ions into CeO2 lattice to form a solid solution results in a
systematic shift of the band at 465cm21 to lower frequencies
with increasing Sn content. Considering that the blue-shift
(red-shift) of Raman F2g peak is associated with lattice expan-
sion (contraction) and weakening (strengthening) of local
cation-anion bonds,58,60 the decreased intensity and the blue
shift of F2g peaks resulting from lattice expansion explicitly
indicate the incorporation of SnO2 into CeO2 fluorite structure
after thermal reduction, with the reduction of CeIV (0.97 Å at
octahedral coordination) ions to bigger CeIII (1.143 Å at octa-
hedral coordination) as supported by the XANES results
below. This is in good accordance with conclusion from afore-
mentioned XRD observations that a continuous reaction
between CeO2 and SnO2 occurs with increasing SnO2 content.
It is noteworthy that after the subsequent H2O splitting step,
the intensity of the F2g peaks increase with a red shift ranging
between 13 to 15 cm21 for cycled samples when x 5 0.05–
0.15 (Figures 5a, b), implying the regeneration of CeO2
fluorite-type structure and lattice contraction with the oxida-
tion of CeIII to smaller CeIV ions. However, for the cycled
sample with x 5 0.20, the F2g peak is neither enhanced in
Figure 5. (a) Raman spectra of CeO2–xSnO2 solid com- intensity nor shifted back to higher wavenumbers, indicating
pounds after various processing steps (as CeO2 and SnO2 are not regenerated completely during reoxi-
synthesized samples calcined at 14008C, dation with H2O. These results are in line with the XRD inves-
samples thermally reduced in flowing Ar, tigations shown in Figure 4c. Meanwhile, cassiterite SnO2,
samples after the subsequent H2O splitting which falls below XRD detection limits (Figure 4c), is identi-
cycle) with different SnO2 content. (b) Plot of fied with characteristic weak Raman bands at ca. 479 cm21
DF2g Raman shift of samples thermally (Supporting Information Figure S3c). This suggests the migra-
reduced as compared to the as synthesized tion of SnO2 out of CeO2 fluorite structure which alleviates
samples (blue shift) and samples after the CeO2 lattice distortions after reoxidation with H2O.
subsequent H2O splitting cycle as compared To provide element-specific electronic configurations of the
to samples thermally reduced (red shift) as a
metals studied and to give better insight into redox chemistry,
function of SnO2 molar ratio.
valence states of Ce and Sn in CeO2–0.15SnO2 solid com-
[Color figure can be viewed at wileyonlinelibrary.com] pound featuring the highest H2 yield were determined by Ce
LIII-edge XANES spectroscopy and 119Sn M€ossbauer spectros-
(Supporting Information Figure S5a).57 The F2g peak origi- copy. Figure 6 shows the normalized XANES spectra for
nates from the symmetrical stretching vibration of the oxygen CeO2–0.15SnO2 after various processing steps together with
atoms located around eight-fold coordination CeIV ions (Sup- the spectra of standard reference compounds at the Ce LIII
porting Information Figure S5d).58–60 Furthermore, weaker edge. XANES spectrum of CeIII(NO3)36H2O (red) is dominat-
Raman bands at ca. 479 cm21, which is the characteristic Eg ed by one intense absorption band at around 5728 eV. The
vibration mode of cassiterite SnO2,61 are observed with spectrum of CeIVO2 (black) exhibits a typical double white
x 5 0.15 and x 5 0.20 (Supporting Information Figure S3a). line 5 eV apart, which stems from the two cerium ground-state
However, this low-intensity band at ca. 479 cm21 is absent electronic configurations of 4f0 and 4f1.66,67 For the initial
with x 5 0.05 and x 5 0.10 (Supporting Information Figure CeO2–0.15SnO2 (green), the spectrum shows exclusively fea-
S3a) attributable to the comparably lower SnO2 content. The tures characteristic of CeIV. Negligible shifts in the edge posi-
above observations clearly confirm that SnO2 has not been tion of the Ce LIII absorption edge is observed as compared to
incorporated into the CeO2 framework of the initial samples, the reference spectra of CeO2. The inset highlights the shift of
as evidenced by the X-ray diffraction patterns in Figure 4a. the absorption edge position of CeO2–0.15SnO2 solid com-
Importantly, all the samples thermally reduced in following pound at three different stages of thermochemical H2O split-
Ar generally decrease remarkably in the intensity of the F2g ting cycling. We note that for CeO2–0.15SnO2 after thermal
peaks, accompanied by a shift to lower wavenumbers with reduction in following Ar (blue), a shift in the edge position to
broader modes (as displayed in Figure 5a). By closer inspec- lower photon energy is evident concomitant with appearance
tion, a continuous blue shift of 25 to 210 cm21 is measured of a new peak corresponding to CeIII (Figure 6) when

3456 DOI 10.1002/aic Published on behalf of the AIChE August 2017 Vol. 63, No. 8 AIChE Journal
s21). This finding implies an evolution from crystalline SnO2
to Ce2Sn2O7 pyrochlore occurs after thermal reduction, corre-
lating well with the XRD results in Figure 4. The relatively
higher IS value suggests an increase in Sn 5s electron density
due to the electron transfer from neighboring Ce to Sn in
Ce2Sn2O7 pyrochlore, while the larger QS value indicating a
more asymmetric environment originating from the surround-
ing Ce atoms as the original Sn neighbors in SnO2 are replaced
with Ce in Ce2Sn2O7 pyrochlore (Supporting Information Fig-
ures S5c, f). Note that no detectable spectroscopic indications

Figure 6. Normalized XANES spectra for CeO2–


0.15SnO2 solid compounds after various
processing steps at the Ce LIII edge togeth-
er with the spectra of standard reference
compounds. Inset represents the shift of
the absorption edge position of CeO2–
0.15SnO2 solid compound as synthesized
(green), reduced (blue), and cycled (brown).
[Color figure can be viewed at wileyonlinelibrary.com]

compared with the as synthesized sample (green), indicating a


reduction from CeIV to CeIII happens. For CeO2–0.15SnO2
samples after the subsequent H2O splitting cycle (brown),
there is a shift in the edge position back to higher photon ener-
gy, which confirms that CeIII reverts back to CeIV during reox-
idation with H2O. Additionally, a slight shift to higher photon
energy compared to the initial spectrum is observed, revealing
the incomplete reoxidation of cerium, which is likely due to
the comparably short reoxidation period (8 min).
119
Sn M€ossbauer spectroscopy is traditionally considered as
the most sensitive tool for probing the oxidation state and the
local structural symmetry of Sn on the atomic scale.68,69 Two
characteristic hyperfine parameters can be extracted from the
measured spectra: isomer shift (IS) and quadrupole splitting
(QS). IS reflects the changes to the s-electrons density, which
is induced by electron-donating or electron-withdrawing
groups and, to a greater extent, by the change of the valence
state. QS indicates the local structural symmetry of the probed
Sn nucleus, with asymmetric environments corresponding to
large splitting of the nuclear energy levels.
The experimental and fitted 119Sn M€ossbauer spectra of
CeO2–0.15SnO2 solid compounds (as synthesized, reduced,
and cycled) are illustrated in Figure 7 and the resulting hyper-
fine parameters of the corresponding fits are included in Sup-
porting Information Table S4. The spectrum of the initial
CeO2–0.15SnO2 (Figure 7a) shows a symmetric Lorentzian Figure 7. Room temperature 119Sn Mo € ssbauer spectra
doublet that can be unambiguously ascribed to typical crystal- of CeO2–0.15SnO2 solid compounds (a) as
line SnO2. The isomer shift close to 0 mm s21 (IS 5 0.088 mm synthesized, (b) thermally reduced in flowing
s21) is characteristic of the SnIV oxidation state and the quad- Ar, and (c) after the subsequent H2O splitting
rupole splitting of QS 5 0.528 mm s21 arises from the intrin- cycle. Structures inset corresponding to the
sic asymmetry of the SnO6 octahedron in the well-crystallized local coordination of (a) SnO2 moiety, (b)
SnO2 (Supporting Information Figures S5b, e). For CeO2– Ce2Sn2O7 moiety, (c) SnO2 and Ce2Sn2O7
0.15SnO2 sample thermally reduced (Figure 7b), the moieties as assigned in literature.
M€ossbauer experimental data can be fitted to a symmetric Lor- The red sphere represents an oxygen atom, whereas
entzian doublet with a higher value of an isomer shift and a green and brown spheres identify tin and cerium atoms,
respectively. [Color figure can be viewed at wileyonline-
quadrupole splitting which can be assigned to SnIV in library.com]
Ce2Sn2O7 pyrochlore (IS 5 0.163 mm s21, QS 5 0.654 mm

AIChE Journal August 2017 Vol. 63, No. 8 Published on behalf of the AIChE DOI 10.1002/aic 3457
Figure 8. Representative SEM images of initial CeO2–xSnO2 samples thermally calcined at 14008C with different
SnO2 content: (a) x 5 0.05, (b) x 5 0.10, (c) x 5 0.15, (d) x 5 0.20, and (e) pure CeO2.

for the reduced Sn species, such as SnII and Sn0, are observed, instead of an innocent spectator, the presence of Sn facilitates
which clearly demonstrates no change in SnIV oxidation state the reversibility of the reduction/oxidation between CeIV/CeIII
occurs for CeO2–0.15SnO2 sample thermally reduced. For for forming the thermodynamically stable Ce2Sn2O7 pyro-
CeO2–0.15SnO2 samples after the subsequent H2O splitting chlore intermediate.
cycle (Figure 7c), the spectrum consists of two symmetric Lor- Structures and morphologies of the as-synthesized and ther-
entzian doublets corresponding to SnO2 and Ce2Sn2O7 with an mochemical cycled samples were studied by scanning electron
area fraction of, respectively, 73% and 27%. This result is microscopy (SEM). Figure 8 shows the typical SEM micro-
well in accordance with the O2 and H2 productivity shown in graphs of initial CeO2–xSnO2 (x 5 0.05–0.20) samples prior to
Supporting Information Table S2 (O2 yield: 4.87 mL g21, H2 thermochemical cycling. For comparison, a SEM image of
yield: 7.18 mL g21), where 74% of the reduced Ce2Sn2O7 was pure CeO2 is given in Figure 8e. Pure CeO2 mainly consists of
reoxidized to produce H2. To sum up, the XANES and small particles with primary diameters of around 0.5 lm.
M€ossbauer spectroscopy results provide unambiguous evi- CeO2–xSnO2 (x 5 0.05, 0.10) exhibits a relatively homoge-
dence that the thermochemical H2O splitting cycling with neous microstructure with smooth surfaces and grain sizes of
CeO2–0.15SnO2 rely on the facile and reversible change of the approximately 1 lm (Figures 8a, b). However, lower homoge-
oxidation state of Ce (CeIVO2 $ CeIII 2 Sn2O7); meanwhile, neity is observed with x 5 0.15 and x 5 0.20 (Figures 8c, d).

3458 DOI 10.1002/aic Published on behalf of the AIChE August 2017 Vol. 63, No. 8 AIChE Journal
Figure 9. (a) High-resolution TEM image of initial CeO2–0.15SnO2 samples thermally calcined at 14008C, along with
the respective FFT patterns characteristics of cubic fluorite CeO2. Representative STEM image and EDX
elemental map of initial CeO2–0.15SnO2 sample thermally calcined at 14008C (b–e), sample thermally
reduced in flowing Ar (f–i), and sample after thermochemical H2O splitting cycling (j–m).
[Color figure can be viewed at wileyonlinelibrary.com]

For x 5 0.15, small particles and large agglomerates appear CeO2–0.15SnO2 solid compound was further analyzed by
mixed with sizes ranging between 1 and 3 lm (Figure 8c). TEM and STEM-EDX. For the initial CeO2–0.15SnO2 sample
Further increasing SnO2 content to x 5 0.20, small grains are thermally calcined at 14008C, the clear lattice fringes observed
almost vanished and only large agglomerates more than 5 lm in high-resolution TEM and the corresponding fast Fourier
are evident (Figure 8d). The BET surface area of the as- transform (FFT) diffraction pattern (Figure 9a) indexed to
synthesized CeO2–xSnO2 (x 5 0.05–0.20) and pure CeO2 are cubic fluorite CeO2 are revealed. However, cassiterite SnO2 is
summarized in Supporting Information Table S5. The lowest not identified, although some variability in contrast across the
surface area of 0.21 m2 g21 was measured for CeO2– investigated region is evident (refer to Supporting Information
0.20SnO2. After thermochemical H2O splitting cycling, parti- Figure S7 for further details). Conversely, STEM-EDX analy-
cle growth is observed over all the samples (Supporting Infor- sis does reveal the presence of CeO2 and SnO2 with Ce and Sn
mation Figure S6). The grain boundaries of all the cycled emissions (Figures 9b–e). Moreover, EDX mapping of Ce and
samples appear to be completely fused. Remarkably, CeO2– Sn displays an inhomogeneous distribution through the as-
0.20SnO2 undergoes dramatic sintering with grains significant- synthesized sample prior to thermochemical cycling, as evi-
ly larger than CeO2–xSnO2 (x 5 0.05–0.15) and pure CeO2. dent by the Ce/Sn segregation phases (Figures 9b-e).

AIChE Journal August 2017 Vol. 63, No. 8 Published on behalf of the AIChE DOI 10.1002/aic 3459
Following thermal reduction in the following Ar, CeO2– SnO2 content. During the subsequent low-temperature reoxi-
0.15SnO2 sample in the reduced state exhibits significant mix- dation with H2O, CeO2, and SnO2 were regenerated, complet-
ing of the CeO2 and SnO2 phases with a homogeneous disper- ing the redox cycle between CeIV and CeIII at 14008C
sion of Ce and Sn elements (Figures 9f–i), further confirming reduction/8008C reoxidation. Encouragingly, H2 yield for
that a reaction between CeO2 and SnO2 occurs after thermal CeO2–0.15SnO2 was improved by a factor of 3.8 due to the
reduction. Additionally, for sample after thermochemical H2O enhanced reduction extent when compared with the state-of-
splitting cycling, Ce and Sn elements are also well mixed (Fig- the-art CeO2, with stable O2 and H2 productivity over seven
ures 9j–m). See Supporting Information Figure S8 for further cycles.
details. In accordance, we can conclude that of the initial It may be anticipated that this novel compound-forming
CeO2–xSnO2 (x 5 0.05–0.20) solid compounds, SnO2 was dis- stoichiometric redox chemistry can be readily implemented
tributed unevenly on CeO2, forming mixed metal oxides. After and exploited with other pyrochlore oxides exhibiting wide
thermal reduction, a reaction between CeO2 and SnO2 occurs, varieties of chemical substitution, thus opening up new oppor-
resulting in a homogeneous dispersion of Ce and Sn elements. tunities for identifying even more promising materials and
accelerating realization of solar fuels production based on
Reaction mechanism for the novel CeO2–xSnO2/
thermochemical cycles.
Ce2Sn2O7 pyrochlore cycle
The experimental findings of two-step thermochemical H2O Acknowledgments
splitting performance and the characterization evidence pre-
sented above provide a basis for refining the mechanism for This work was supported by National Natural Science
the novel CeO2–xSnO2/Ce2Sn2O7 pyrochlore cycle. We con- Foundation of China (21476232, 21573232, 21522608,
clude that instead of proceeding via the nonstoichiometric O- 21676269) and the “Strategic Priority Research Program” of
vacancy mechanism for pure CeO2 as embodied by Eqs. 3 and the Chinese Academy of Sciences (XDB17020100).
4, a new CeO2–xSnO2/Ce2Sn2O7 pyrochlore stoichiometric
Literature Cited
redox cycle is involved for CeO2–xSnO2 in two-step thermo-
chemical H2O splitting. This new redox cycle is illustrated as 1. Lewis NS, Nocera DG. Powering the planet: chemical challenges in
solar energy utilization. Proc Natl Acad Sci USA. 2006;103:15729–
follows: 15735.
High-temperature reduction 2. Trainham JA, Newman J, Bonino CA, Hoertz PG, Akunuri N.
1 1 Whither solar fuels? Curr Opin Chem Eng. 2012;1:204–210.
CeIV O2 1SnO2 ! CeIII 2 Sn2 O7 1 O2 ðgÞ (6) 3. Centi G, Perathoner S. Towards solar fuels from water and CO2.
2 4 ChemSusChem. 2010;3:195–208.
4. Liu J, Liu Y, Liu N, Han Y, Zhang X, Huang H, Lifshitz Y, Lee S-
Low-temperature reoxidation with H2O T, Zhong J, Kang Z. Metal-free efficient photocatalyst for stable vis-
1 III 1 1 ible water splitting via a two-electron pathway. Science. 2015;347:
Ce 2 Sn2 O7 1 H2 OðgÞ ! CeIV O2 1 SnO2 1 H2 ðgÞ (7) 970–974.
2 2 2 5. Xiang Q, Cheng B, Yu J. Graphene-based photocatalysts for solar-
To summarize, during the first high-temperature reduction fuel generation. Angew Chem Int Ed. 2015;54:11350–11366.
6. Luo J, Im J-H, Mayer MT, Schreier M, Nazeeruddin MK, Park N-G,
step (14008C), a stoichiometric reaction between inhomoge- Tilley SD, Fan HJ, Graetzel M. Water photolysis at 12.3% efficiency
neous distributed CeO2 and SnO2 occurred, forming thermo- via perovskite photovoltaics and earth-abundant catalysts. Science.
dynamically stable Ce2Sn2O7 pyrochlore instead of metastable 2014;345:1593–1596.
CeO2-d (Eq. 6). As a result, enhanced reduction extent was 7. Cox CR, Lee JZ, Nocera DG, Buonassisi T. Ten-percent solar-to-
fuel conversion with nonprecious materials. Proc Natl Acad Sci
accomplished on account of the reduction of CeIV to CeIII.
USA. 2014;111:14057–14061.
During the subsequent low-temperature reoxidation with H2O 8. Chueh WC, Falter C, Abbott M, Scipio D, Furler P, Haile SM,
(8008C), the reduced Ce2Sn2O7 pyrochlore was reoxidized Steinfeld A. High-flux solar-driven thermochemical dissociation of
with CeIII reverted back to CeIV. Meanwhile CeO2 and SnO2 CO2 and H2O using nonstoichiometric ceria. Science. 2010;330:
were regenerated with a homogeneous dispersion, completing 1797–1801.
9. Muhich CL, Evanko BW, Weston KC, Lichty P, Liang X, Martinek
the redox cycle between CeIV and CeIII at modest temperatures J, Musgrave CB, Weimer AW. Efficient generation of H2 by split-
(Eq. 7). ting water with an isothermal redox cycle. Science. 2013;341:540–
542.
Conclusions 10. Roeb M, Sattler C. Isothermal water splitting. Science. 2013;341:
470–471.
In this contribution, two-step STWS operating via a stoi- 11. Meng X, Wang T, Liu L, Ouyang S, Li P, Hu H, Kako T, Iwai H,
chiometric reaction mechanism was successfully demonstrated Tanaka A, Ye J. Photothermal conversion of CO2 into CH4 with H2
with the successive insertion/extraction of SnO2 within the over group VIII nanocatalysts: an alternative approach for solar fuel
production. Angew Chem Int Ed. 2014;53:11478–11482.
intriguing CeO2–xSnO2/Ce2Sn2O7 pyrochlore cycle. The nov- 12. Chanmanee W, Islam MF, Dennis BH, MacDonnell FM. Solar pho-
el pyrochlore cycle was corroborated using X-ray diffraction, tothermochemical alkane reverse combustion. Proc Natl Acad Sci
Raman spectroscopy, XANES spectroscopy, M€ossbauer spec- USA. 2016;113:2579–2584.
troscopy, in combination with electron microscopy (SEM, 13. Fletcher EA, Moen RL. Hydrogen and oxygen from water. Science.
1977;197:1050–1056.
HRTEM, STEM-EDX). The first thermal reduction proceeded 14. Funk JE. Thermochemical hydrogen production: past and present.
via a stoichiometric reaction between CeO2 and SnO2, forming Int J Hydrogen Energy. 2001;26:185–190.
thermodynamically stable Ce2Sn2O7 pyrochlore rather than 15. Kodama T, Gokon N. Thermochemical cycles for high-temperature
metastable CeO2-d. Consequently, enhanced reduction extents solar hydrogen production. Chem Rev. 2007;107:4048–4077.
were achieved for all the CeO2–xSnO2 (x 5 0.05–0.20) solid 16. Perkins C, Weimer AW. Solar-thermal production of renewable
hydrogen. AIChE J. 2009;55:286–293.
compounds owing to the reduction of CeIV to CeIII. O2 produc- 17. Arifin D, Aston VJ, Liang X, McDaniel AH, Weimer AW. CoFe2O4
tion capacities were improved significantly when compared on a porous Al2O3 nanostructure for solar thermochemical CO2 split-
with pure CeO2, which increased linearly with increasing ting. Energy Environ Sci. 2012;5:9438.

3460 DOI 10.1002/aic Published on behalf of the AIChE August 2017 Vol. 63, No. 8 AIChE Journal
18. Scheffe JR, McDaniel AH, Allendorf MD, Weimer AW. Kinetics and and carbon dioxide by solar-driven thermochemical cycles using rho-
mechanism of solar-thermochemical H2 production by oxidation of a dium-ceria. Energy Environ Sci. 2016;9:2400–2409.
cobalt ferrite–zirconia composite. Energy Environ Sci. 2013;6:963. 41. Abanades S, Legal A, Cordier A, Peraudeau G, Flamant G, Julbe A.
19. Muhich CL, Ehrhart BD, Witte VA, Miller SL, Coker EN, Investigation of reactive cerium-based oxides for H2 production by ther-
Musgrave CB, Weimer AW. Predicting the solar thermochemical mochemical two-step water-splitting. J Mater Sci. 2010;45:4163–4173.
water splitting ability and reaction mechanism of metal oxides: a 42. Scheffe JR, Jacot R, Patzke GR, Steinfeld A. Synthesis, characterization,
case study of the hercynite family of water splitting cycles. Energy and thermochemical redox performance of Hf41, Zr41, and Sc31 doped
Environ Sci. 2015;8:3687–3699. ceria for splitting CO2. J Phys Chem C. 2013;117:24104–24114.
20. McDaniel AH, Miller EC, Arifin D, Ambrosini A, Coker EN, O’hayre 43. Hao Y, Yang C-K, Haile SM. Ceria–zirconia solid solutions (Ce1–
R, Chueh WC, Tong J. Sr- and Mn-doped LaAlO32d for solar thermo- xZrxO22d, x 0.2) for solar thermochemical water splitting: a ther-
chemical H2 and CO Production. Energy Environ Sci. 2013;6:2424. modynamic study. Chem Mater. 2014;26:6073–6082.
21. Furler P, Scheffe JR, Steinfeld A. Syngas production by simulta- 44. Rothensteiner M, Bonk A, Vogt UF, Emerich H, van Bokhoven JA.
neous splitting of H2O and CO2 via ceria redox reactions in a high- Structural changes in Ce0.5Zr0.5O22d under temperature-swing and
temperature solar reactor. Energy Environ Sci. 2012;5:6098–6103. isothermal solar thermochemical looping conditions determined by
22. Ohya H, Yatabe M, Aihara M, Negishi Y, Takeuchi T. Feasibility of in situ Ce K and Zr K edge X-ray absorption spectroscopy. J Phys
hydrogen production above 2500 K by direct thermal decomposition Chem C. 2016;120:13931–13941.
reaction in membrane reactor using solar energy. Int J Hydrogen 45. Petkovich ND, Rudisill SG, Venstrom LJ, Boman DB, Davidson JH,
Energy. 2002;27:369–376. Stein A. Control of heterogeneity in nanostructured Ce1–xZrxO2 bina-
23. Loutzenhiser PG, Meier A, Steinfeld A. Review of the two-step ry oxides for enhanced thermal stability and water splitting activity.
H2O/CO2-splitting solar thermochemical cycle based on Zn/ZnO J Phys Chem C. 2011;115:21022–21033.
redox reactions. Materials. 2010;3:4922–4938. 46. Tolla B, Demourgues A, Isnard O, Menetrier M, Pouchard M,
24. Scheffe JR, Steinfeld A. Oxygen exchange materials for solar ther- Rabardel L, Seguelong T. Structural investigation of oxygen inser-
mochemical splitting of H2O and CO2: a review. Mater Today. tion within the Ce2Sn2O7-Ce2Sn2O8 pyrochlore solid solution by
2014;17:341–348. means of in situ neutron diffraction experiments. J Mater Chem.
25. Muhich CL, Ehrhart BD, Al-Shankiti I, Ward BJ, Musgrave CB, 1999;9:3131–3136.
Weimer AW. A review and perspective of efficient hydrogen genera- 47. Subramanian MA, Aravamudan G, Rao GVS. Oxide pyrochlores2a
tion via solar thermal water splitting. Wiley Interdiscip Rev Energy review. Prog Solid State Chem. 1983;15:55–143.
Environ. 2016;5:261–287. 48. Thomson JB, Armstrong AR, Bruce PG. A new class of pyrochlore
26. Gokon N, Hasegawa T, Takahashi S, Kodama T. Thermochemical solid solution formed by chemical intercalation of oxygen. J Am
two-step water-splitting for hydrogen production using Fe-YSZ par- Chem Soc. 1996;118:11129–11133.
ticles and a ceramic foam device. Energy. 2008;33:1407–1416. 49. Sickafus KE, Mineevini L, Grimes RE, Valde JA, Ishimare M, Li F,
27. Gokon N, Murayama H, Umeda J, Hatamachi T, Kodama T. Mono- McClellan KJ, Hartmann T. Radiation tolerance of complex oxides.
clinic zirconia-supported Fe3O4 for the two-step water-splitting ther- Science. 2000;289:748–751.
mochemical cycle at high thermal reduction temperatures of 1400– 50. Wright CS, Fisher J, Thompsett D, Walton RI. Hydrothermal synthe-
16008C. Int J Hydrogen Energy. 2009;34:1208–1217. sis of a cerium(IV) pyrochlore with low-temperature redox proper-
28. Stamatiou A, Loutzenhiser PG, Steinfeld A. Solar syngas production ties. Angew Chem Int Ed. 2006;45:2442–2446.
via H2O/CO2-splitting thermochemical cycles with Zn/ZnO and FeO/ 51. Yamamoto T, Suzuki A, Nagai Y, Tanabe T, Dong F, Inada Y,
Fe3O4 redox reactions. Chem Mater. 2010;22:851–859. Nomura M, Tada M, Iwasawa Y. Origin and dynamics of oxygen
29. Stamatiou A, Loutzenhiser PG, Steinfeld A. Syngas production from storage/release in a Pt/ordered CeO2–ZrO2 catalyst studied by time-
H2O and CO2 over Zn particles in a packed-bed reactor. AIChE J. resolved XAFS analysis. Angew Chem Int Ed. 2007;119:9413–9416.
2012;58:625–631. 52. Sickafus KE, Grimes RW, Valdez JA, Cleave A, Tang M, Ishimaru
30. Charvin P, Abanades S, Lemont F, Flamant G. Experimental study M, Corish SM, Stanek CR, Uberuaga BP. Radiation-induced
of SnO2/SnO/Sn thermochemical systems for solar production of amorphization resistance and radiation tolerance in structurally relat-
hydrogen. AIChE J. 2008;54:2759–2767. ed oxides. Nat Mater. 2007;6:217–223.
31. Lev^eque G, Abanades S, Jumas J-C, Olivier-Fourcade J. Characteri- 53. Oh SH, Black R, Pomerantseva E, Lee JH, Nazar LF. Synthesis of a
zation of two-step tin-based redox system for thermochemical fuel metallic mesoporous pyrochlore as a catalyst for lithium-O2 batter-
production from solar-driven CO2 and H2O splitting cycle. Ind Eng ies. Nat Chem. 2012;4:1004–1010.
Chem Res. 2014;53:5668–5677. 54. Shamblin J, Feygenson M, Neuefeind J, Tracy CL, Zhang F,
32. Chueh WC, Haile SM. Ceria as a thermochemical reaction medium Finkeldei S, Bosbach D, Zhou H, Ewing RC, Lang M. Probing dis-
for selectively generating syngas or methane from H2O and CO2. order in isometric pyrochlore and related complex oxides. Nat
ChemSusChem. 2009;2:735–739. Mater. 2016;15:507–511.
33. Chueh WC, Haile SM. A Thermochemical study of ceria: exploiting 55. Scheffe JR, Allendorf MD, Coker EN, Jacobs BW, McDaniel AH,
an old material for new modes of energy conversion and CO2 miti- Weimer AW. Hydrogen production via chemical looping redox
gation. Philos Trans R Soc A. 2010;368:3269–3294. cycles using atomic layer deposition-synthesized iron oxide and
34. Marxer D, Furler P, Scheffe J, Geerlings H, Falter C, Batteiger V, cobalt ferrites. Chem Mater. 2011;23:2030–2038.
Sizmann A, Steinfeld A. Demonstration of the entire production 56. Tolla B, Demourgues A, Pouchard M, Rabardel L, Fournes L,
chain to renewable kerosene via solar thermochemical splitting of Wattiaux A. Oxygen exchange properties in the new pyrochlore solid
H2O and CO2. Energy Fuels. 2015;29:3241–3250. solution Ce2Sn2O7-Ce2Sn2O8. Solid State Chem Cryst Chem. 1999;2:
35. Abanades S, Flamant G. Thermochemical hydrogen production from 139–146.
a two-step solar-driven water-splitting cycle based on cerium oxides. 57. Guzman J, Carrettin S, Corma A. Spectroscopic evidence for the supply
Sol Energy. 2006;80:1611–1623. of reactive oxygen during CO oxidation catalyzed by gold supported on
36. Meng Q-L, Lee C-I, Ishihara T, Kaneko H, Tamaura Y. Reactivity nanocrystalline CeO2. J Am Chem Soc. 2005;127:3286–3287.
of CeO2-based ceramics for solar hydrogen production via a two- 58. McBride JR, Hass KC, Poindexter BD, Weber WH. Raman and X-
step water-splitting cycle with concentrated solar energy. Int J ray studies of Ce12xRexO22y, where Re5La, Pr, Nd, Eu, Gd, and
Hydrogen Energy. 2011;36:13435–13441. Tb. J Appl Phys. 1994;76:2435.
37. Scheffe JR, Steinfeld A. Thermodynamic analysis of cerium-based 59. Banerji A, Grover V, Sathe V, Deb SK, Tyagi AK. CeO2-Gd2O3
oxides for solar thermochemical fuel production. Energy Fuels. system: unraveling of microscopic features by Raman spectroscopy.
2012;26:1928–1936. Solid State Commun. 2009;149:1689–1692.
38. Kaneko H, Ishihara H, Taku S, Naganuma Y, Hasegawa N, Tamaura 60. Rupp JLM, Fabbri E, Marrocchelli D, Han J-W, Chen D, Traversa
Y. Cerium ion redox system in CeO2 – xFe2O3 solid solution at high E, Tuller HL, Yildiz B. Scalable oxygen-ion transport kinetics in
temperatures (1,273–1,673 k) in the two-step water-splitting reaction metal-oxide films: impact of thermally induced lattice compaction in
for solar H2 generation. J Mater Sci. 2008;43:3153–3161. acceptor doped ceria films. Adv Funct Mater. 2014;24:1562–1574.
39. Call F, Roeb M, Schm€ ucker M, Sattler C, Pitz-Paal R. Ceria doped 61. Johari A, Bhatnagar MC, Rana V. Effect of substrates on structural
with zirconium and lanthanide oxides to enhance solar thermochemi- and optical properties of tin oxide (SnO2) nanostructures. J Nanosci
cal production of fuels. J Phys Chem C. 2015;119:6929–6938. Nanotechnol. 2012;12:7903–7908.
40. Lin FJ, Rothensteiner M, Alxneit I, van Bokhoven JA, Wokaun A. 62. Yang N, Shi Y, Schweiger S, Strelcov E, Belianinov A, Foglietti V,
First demonstration of direct hydrocarbon fuel production from water Orgiani P, Balestrino G, Kalinin SV, Rupp JLM, Aruta C. Role of

AIChE Journal August 2017 Vol. 63, No. 8 Published on behalf of the AIChE DOI 10.1002/aic 3461
associated defects in oxygen ion conduction and surface exchange 67. Zhang F, Wang P, Koberstein J, Khalid S, Chan S-W. Cerium oxida-
reaction for epitaxial Samaria-doped ceria thin films as catalytic tion state in ceria nanoparticles studied with X-ray photoelectron
coatings. ACS Appl Mater Interfaces. 2016;8:14613–14621. spectroscopy and absorption near edge spectroscopy. Surf Sci. 2004;
63. Giannici F, Gregori G, Aliotta C, Longo A, Maier J, Martorana A. Structure 563:74–82.
and oxide ion conductivity: local order, defect interactions and grain bound- 68. de Kergommeaux A, Faure-Vincent J, Pron A, de Bettignies R,
ary effects in acceptor-doped ceria. Chem Mater. 2014;26:5994–6006. Malaman B, Reiss P. Surface oxidation of tin chalcogenide nano-
64. Shi Y, Bork AH, Schweiger S, Rupp JLM. The effect of mechanical crystals revealed by 119Sn-M€ossbauer spectroscopy. J Am Chem Soc.
twisting on oxygen ionic transport in solid-state energy conversion 2012;134:11659–11666.
membranes. Nat Mater. 2015;14:721–727. 69. Protesescu L, Rossini AJ, Kriegner D, Valla M, de Kergommeaux
65. Li LL, Xu J, Yuan Q, Li ZX, Song WG, Yan CH. Facile synthesis A, Walter M, Kravchyk KV, Nachtegaal M, Stangl J, Malaman B,
of macrocellular mesoporous foamlike Ce-Sn mixed oxides with a Reiss P, Lesage A, Emsley L, Coperet C, Kovalenko MV. Unrav-
nanocrystalline framework by using triblock copolymer as the single eling the core-shell structure of ligand-capped Sn/SnOx nanopar-
template. Small. 2009;5:2730–2737. ticles by surface-enhanced nuclear magnetic resonance,
66. Elfallah J, Boujana S, Dexpert H, Kiennemann A, Majerus J, Touret M€ ossbauer, and X-ray absorption spectroscopies. Acs Nano. 2014;
O, Villain F, Lenormand F. Redox processes on pure ceria and on 8:2639–2648.
Rh/CeO2 catalyst monitored by X-ray absorption (Fast acquisition
mode). J Phys Chem. 1994;98:5522–5533. Manuscript received Nov. 16, 2016, and revision received Feb. 13, 2017.

3462 DOI 10.1002/aic Published on behalf of the AIChE August 2017 Vol. 63, No. 8 AIChE Journal

You might also like