ArticlePDF Available

Solar-driven co-thermolysis of CO 2 and H 2 O and in-situ oxygen removal across a non-stoichiometric ceria membrane

Authors:
  • Institute of Engineering Thermophysics

Abstract

We report on the first-ever experimental demonstration of simultaneous thermolysis of CO2 and H2O with in-situ separation of fuel and oxygen in a solar-driven membrane reactor. Gaseous CO2/H2O mixtures at molar ratios from 3:4 to 2:1 were fed to a mixed ionic-electronic conducting non-stoichiometric ceria (CeO2-δ) membrane enclosed in a solar cavity-receiver and exposed to simulated concentrated solar radiation of up to 4200 suns. Reaction rates were measured at isothermal and isobaric conditions in the range of 1723-1873 K and 0.2-1.7 Pa O2, yielding a maximum combined CO and H2 fuel production rate of 2.3 µmol cm⁻² min⁻¹ at 1873 K and 0.2 Pa O2 at steady state, which corresponded to a conversion of reactants of 0.7%. At all conditions tested, CO production was favored over H2 production, as expected from theory. Experimental results followed the same trends as the thermodynamic equilibrium limits of membrane-assisted thermochemical fuel production.
This is an Accepted Manuscript, which has been through the
Royal Society of Chemistry peer review process and has been
accepted for publication.
Accepted Manuscripts are published online shortly after
acceptance, before technical editing, formatting and proof reading.
Using this free service, authors can make their results available
to the community, in citable form, before we publish the edited
article. We will replace this Accepted Manuscript with the edited
and formatted Advance Article as soon as it is available.
You can find more information about Accepted Manuscripts in the
author guidelines.
Please note that technical editing may introduce minor changes
to the text and/or graphics, which may alter content. The journal’s
standard Terms & Conditions and the ethical guidelines, outlined
in our author and reviewer resource centre, still apply. In no
event shall the Royal Society of Chemistry be held responsible
for any errors or omissions in this Accepted Manuscript or any
consequences arising from the use of any information it contains.
Accepted Manuscript
rsc.li/reaction-engineering
ISSN 2058-9883
http://rsc.li/reaction-engineering
Reaction Chemistry
& Engine ering
Linking fundamental chemistry and engineering to create scalable, efficient processes
PAPER
Ryan L. Hartman et al.
Influence of water on the deprotonation and the ionic mechanisms of a
Heck alkynylation and its resultant E-factors
Volume 1 Number 1 1 February 2016 Pages 1–120
Rea c tion Chemis try
& Engineering
View Article Online
View Journal
This article can be cited before page numbers have been issued, to do this please use: M. Tou, J. Jin, Y.
Hao, A. Steinfeld and R. Michalsky, React. Chem. Eng., 2019, DOI: 10.1039/C8RE00218E.
1
Solar-driven co-thermolysis of CO2 and H2O and in-situ oxygen removal
across a non-stoichiometric ceria membrane
Maria Tou1, Jian Jin2,3, Yong Hao2,3, Aldo Steinfeld1, Ronald Michalsky1*
1 Department of Mechanical and Process Engineering, ETH Zürich, 8092 Zürich, Switzerland
2 Institute of Engineering Thermophysics, Chinese Academy of Sciences, 11 Beisihuanxi Rd.,
Beijing 100190, P. R. China
3 University of Chinese Academy of Sciences, No.19A Yuquan Rd., Beijing 100049, P. R. China
*corresponding author
Abstract
We report on the first-ever experimental demonstration of simultaneous thermolysis of CO2 and
H2O with in-situ separation of fuel and oxygen in a solar-driven membrane reactor. Gaseous
CO2/H2O mixtures at molar ratios from 3:4 to 2:1 were fed to a mixed ionic-electronic
conducting non-stoichiometric ceria (CeO2-δ) membrane enclosed in a solar cavity-receiver and
exposed to simulated concentrated solar radiation of up to 4200 suns. Reaction rates were
measured at isothermal and isobaric conditions in the range of 1723-1873 K and 0.2-1.7 Pa O2,
yielding a maximum combined CO and H2 fuel production rate of 2.3 µmol cm-2 min-1 at 1873 K
and 0.2 Pa O2 at steady state, which corresponded to a conversion of reactants of 0.7%. At all
conditions tested, CO production was favored over H2 production, as expected from theory.
Experimental results followed the same trends as the thermodynamic equilibrium limits of
membrane-assisted thermochemical fuel production.
Page 1 of 21 Reaction Chemistry & Engineering
Reaction Chemistry & Engineering Accepted Manuscript
Open Access Article. Published on 14 March 2019. Downloaded on 4/14/2019 7:04:04 PM.
This article is licensed under a
Creative Commons Attribution 3.0 Unported Licence.
View Article Online
DOI: 10.1039/C8RE00218E
2
Introduction
The utilization of the vast solar energy resource for electricity, heat, and fuels has become a key
objective in research and development.1 The conversion and storage of solar energy in fuels is
especially appealing as a means to transition from fossil fuels to a “CO2 economy”.2 For this, a
solar refinery needs to be developed where solar energy is collected and used to convert CO2 and
H2O to fuels by some method. Existing research spans solar-driven electrochemical, photo-
electrochemical, and photocatalytic paths for direct conversion, as well as indirect routes via the
solar thermochemical production of syngas (H2 and CO).3
Solar thermochemical redox cycles utilize the entire spectrum of solar radiation concentrated to
high-temperature process heat to drive the splitting of CO2 and H2O and produce CO and H2 at
high rates, selectivity, mass conversions, and efficiencies.4-6 However, the temperature swing
required between the redox steps induce significant material stresses and energy irreversibilities,
which prompted the search for alternative isothermal processes.7-9 One promising approach is the
use of a dense, ceramic, mixed ionic-electronic conducting (MIEC) membrane for the continuous
separation of oxygen and fuel (H2 and/or CO) derived from the thermolysis of CO2 and H2O at
high temperatures, as pioneered for solar water splitting by Fletcher and co-workers.10, 11 We
recently demonstrated the proof-of-concept utilizing a solar-driven membrane reactor for
splitting of CO2.12 Other investigations of thermochemical membrane reactors, both theoretical
and experimental, have also only focused on either CO2- or H2O-splitting.10, 13-18 This work goes
further and demonstrates the feasibility of co-feeding both CO2 and H2O and assesses the relative
favorability between the two thermolysis reactions occurring simultaneously. The desired
dissociations are chemical equilibrium reactions in the gas phase described by:
(1)
12
2 2
CO CO + O
(2)
12
2 2 2
H O H + O
The reactions are analogous; that is, both are endothermic and thermolytic, but their reaction
energetics differ. This is described by the standard Gibbs free energy changes at equilibrium
G=0):
(3)
2
2
121
( )
CO O 2
1 1
CO
ln ln
p p
G RT p RT K
p
 
 
 
 
 
o
o
Page 2 of 21Reaction Chemistry & Engineering
Reaction Chemistry & Engineering Accepted Manuscript
Open Access Article. Published on 14 March 2019. Downloaded on 4/14/2019 7:04:04 PM.
This article is licensed under a
Creative Commons Attribution 3.0 Unported Licence.
View Article Online
DOI: 10.1039/C8RE00218E
3
(4)
2 2
2
121
( )
H O 2
2 2
ln ln
H O
p p
G RT p RT K
p
 
 
 
 
 
o
o
where ΔGi° is related to the equilibrium constant Ki which in turn is a function of temperature
only. Figure 1a shows T- pO2 equilibrium contours of the separate thermolysis of CO2 and H2O
for various mole fractions of CO or H2 in the product gas, respectively. For both CO2 and H2O
thermolysis, products are favored with increasing T and decreasing pO2, i.e. higher mole fractions
of fuel are possible towards the upper-left corner of Fig. 1a. At such high T, dissociation of CO2
is more thermodynamically favorable than that of H2O at equal conditions.19, 20 A decrease in pO2
can be achieved without use of high-value electrical energy by removal of O2 utilizing a dense
membrane made of an oxygen-selective MIEC material.21 The pO2 is controlled to a low value on
the opposite side of the membrane. For each of the separate thermolysis reactions, if pO2 ½ pCO
or pO2 ½ pH2, the membrane provides no benefit.
The membrane reactor concept used in this work for the co-thermolysis of H2O and CO2 is
shown schematically in Figure 1b. CO2 and H2O are fed to the inner side of a capped tubular
non-stoichiometric ceria membrane. Ceria has become the benchmark material for oxygen-
cycling applications due to its stability and fast kinetics.22-24 In our previous work it was also
found to be an effective material for oxygen-conducting membranes.12 The supply of
concentrated solar process heat at high temperatures drives the thermolysis, producing CO, H2,
and O2. The latter adsorbs at the inner membrane surface, dissociates, and is transported across
the membrane in an ionic form along a chemical potential gradient. The O2- then associates into
O2 at the outer membrane surface and desorbs into an inert sweep gas contained in a shell tube.
This in-situ removal of one of the reaction products drives the reactions forward towards
dissociation and avoids downstream recombination. The counter-flow configuration of the
reactant and product gases favorably maximizes the gradient of pO2 along the length of the
membrane. By placing this reactor in a solar cavity-receiver, the high-temperature heat for the
reactions is provided by concentrated solar radiation incident on the shell tube.
Page 3 of 21 Reaction Chemistry & Engineering
Reaction Chemistry & Engineering Accepted Manuscript
Open Access Article. Published on 14 March 2019. Downloaded on 4/14/2019 7:04:04 PM.
This article is licensed under a
Creative Commons Attribution 3.0 Unported Licence.
View Article Online
DOI: 10.1039/C8RE00218E
4
A B
Fig. 1 (a) Equilibrium contours for separate thermolysis of CO2 (light-colored) or H2O (dark-
colored) at 1 bar total pressure as a function of T and pO2 (according to Eqs. 3 and 4) for various
mole fractions of CO or H2, respectively. Contours extend until pCO or pH2 = 2· pO2, increasing
pO2 beyond this point no longer provides benefit over unperturbed thermolysis. (b) Schematic of
the tubular redox membrane reactor for splitting of CO2 and H2O. CO2 and H2O are fed to the
inner side of the membrane and dissociate into fuel and O2, with the latter selectively crossing
the membrane into Ar sweep gas.
Typical operating conditions demand temperatures around 1773 K and partial pressures of O2
down to 1 Pa. These high temperatures eliminate the need for catalysts but pose significant
constraints on the construction materials which must withstand these conditions over extended
periods of time. Materials must also resist thermal shock that may occur due to cooling
overnight, unless the reactor is equipped with an alternative heat source such as a high-
temperature thermal energy storage system.25 A modular tubular membrane design could avoid
costly maintenance by allowing for simple replacement of degraded membranes. Maintaining
low partial pressures of oxygen is crucial, requiring additional energy for vacuum pumping or
gas separation to regenerate the inert sweep gas (such as N2, though here we use Ar for gas
analytic considerations).26, 27 Alternatively, some studies have reported solar-driven production
Page 4 of 21Reaction Chemistry & Engineering
Reaction Chemistry & Engineering Accepted Manuscript
Open Access Article. Published on 14 March 2019. Downloaded on 4/14/2019 7:04:04 PM.
This article is licensed under a
Creative Commons Attribution 3.0 Unported Licence.
View Article Online
DOI: 10.1039/C8RE00218E
5
of pure O2 and inert gas with low partial pressures of oxygen using thermochemical oxygen
pumps driven by low-grade process heat.27-29
Experimental
Materials. Cerium (IV) oxide (ceria, CeO2, powder, particle size <5 μm, 99.9% purity), poly (oxy-
1,4-phenylene sulfonyl-1,4-phenylene) (PES, (C12H8O3S)n, pellets), polyvinylpyrrolidone (PVP,
(C6H9NO)n, powder, average M.W. 40,000), and 1-methyl-2-pyrrolidinone (NMP, C5H9NO,
liquid, ≥99.0% purity) were from Sigma Aldrich. Al2O3 membranes (Alsint 99.7, 7 mm outer
diameter (OD), 5 mm inner diameter (ID), 250 mm length) were from Intertechno-Firag AG. High-
purity alumina adhesive (Aremco Ceramabond 569) and glass-filled sealant (Aremco-Seal 617)
were from Kager Industrieprodukte GmbH. CO2 (99.998%), Ar (99.996%, 99.999%), He
(99.999%), and calibration gas mixtures, i.e., 1000 mol ppm H2 (99.999%) and 500 mol ppm CO
(99.997%) in Ar (99.999%), and 1000 mol ppm CO (99.997%), 500 mol ppm CO2 (99.995%), 500
mol ppm N2 (99.999%), and 100 mol ppm O2 (99.999%) in Ar (99.999%) were from Messer
Schweiz AG. According to the manufacturer, Ar (99.996%) contained < 5 ppm O2 on volume
basis, equivalent to a limiting pO2 < 0.5 Pa.
Membrane Fabrication. Capped tubular ceria membranes were produced as reported previously.12
Briefly, membranes were fabricated using a phase-inversion/sintering method.30, 31 Two polymers,
namely PES (5.7 wt%) and PVP (0.5 wt%), were dissolved in NMP (22.0 wt%). Ceria powder
(71.8 wt%) was suspended in the polymer solution. The ceria slurry was coated onto membrane
templates (High-Flexible silicone tubing, 3 mm ID, 7 mm OD, RCT Reichelt Chemietechnik
GMbH & Co.), which were placed into a water bath for phase inversion (unfiltered tap water
coagulant at ambient conditions). The silicone templates were removed, and the dried membrane
precursors were then sintered for 8 hours at 1873 K (oven model HTL 20/17, ThermConcept). The
sintered membranes were 6-7 mm OD, 5-6 mm ID, and 150-250 mm in length. Typically, the
membrane walls were about 0.5 mm thick. In-depth solid-state characterization of ceria
membranes before and after use in the reactor was performed previously.12 SEM analysis in Fig.
S1 shows that ceria membranes exposed to both CO2 and H2O in thermolysis experiments do not
change morphologically, consistent with membranes used in pure-CO2 experiments.
Page 5 of 21 Reaction Chemistry & Engineering
Reaction Chemistry & Engineering Accepted Manuscript
Open Access Article. Published on 14 March 2019. Downloaded on 4/14/2019 7:04:04 PM.
This article is licensed under a
Creative Commons Attribution 3.0 Unported Licence.
View Article Online
DOI: 10.1039/C8RE00218E
6
Experimental setup. The solar membrane reactor system is depicted schematically in Figure 2. CO2
and/or H2O was fed through a feeder tube into the inner side of the membrane while Ar sweep gas
was fed into the reactor shell tube in a counter-current flow. This assembly was placed in a
thermally insulated solar cavity-receiver with an aperture 4 cm in diameter. A compound parabolic
concentrator (CPC) was incorporated onto the aperture to boost the solar flux concentration and
generate a more uniform directional distribution of concentrated radiation entering the cavity.32
Experimentation was performed at the High-Flux Solar Simulator (HFSS) of ETH Zurich: an array
of seven Xe arcs, close-coupled to truncated ellipsoidal reflectors, provided an external source of
intense thermal radiation that closely approximated the heat transfer characteristics of highly
concentrating solar energy facilities. The radiative flux distribution at the focal plane was
measured optically using a calibrated CCD camera focused on a Lambertian (diffusely reflecting)
target. The solar radiative power input to the cavity was calculated by integration of the radiative
flux over the aperture area and verified with a water calorimeter. Temperatures were measured at
the outer surface of the reactor shell at two heights along the tube (indicated in Figure 2) using B-
type thermocouples. Gas flow rates were regulated by electronic mass flow controllers (MFC,
Bronkhorst F-201 C, accuracy 0.5%Rd + 0.1%FS), whereas steam flow was generated by a liquid
flow controller (LFC, Bronkhorst Liqui-Flow L23-AAD-33-K-305, accuracy 1%FS) and steam
generator (Bronkhorst CEM W-202A-333-K). Product gas composition in each stream was
monitored on-line by gas chromatography (GC, Agilent 490 MicroGC).
Page 6 of 21Reaction Chemistry & Engineering
Reaction Chemistry & Engineering Accepted Manuscript
Open Access Article. Published on 14 March 2019. Downloaded on 4/14/2019 7:04:04 PM.
This article is licensed under a
Creative Commons Attribution 3.0 Unported Licence.
View Article Online
DOI: 10.1039/C8RE00218E
7
Fig. 2 Schematic and flow diagram of the experimental setup used to test co-thermolysis of CO2
and H2O in a membrane reactor. Simulated concentrated solar radiation from the HFSS enters the
cavity-receiver through the aperture and heats the reactor. Gaseous CO2 and/or H2O flows through
a feeder tube into the membrane, then flows upward through the annulus between the membrane
Page 7 of 21 Reaction Chemistry & Engineering
Reaction Chemistry & Engineering Accepted Manuscript
Open Access Article. Published on 14 March 2019. Downloaded on 4/14/2019 7:04:04 PM.
This article is licensed under a
Creative Commons Attribution 3.0 Unported Licence.
View Article Online
DOI: 10.1039/C8RE00218E
8
and feeder tubes before exiting the reactor. The membrane removes O2 produced from thermolysis.
Unreacted H2O is removed in the condenser and the composition of the remaining gas is analyzed
by GC1. In counter-current to the oxidant stream, a sweep gas (Ar) flows through the shell tube,
taking up O2 crossing the membrane, and exits below the cavity-receiver to be analyzed by GC2.
Not to scale.
Experimental runs. All volumetric flow rates are given at standard conditions (1 bar and 273 K).
The HFSS heated the reactor to the desired temperatures in the range 1723-1873 K with a radiative
power input of 2.5-3.0 kW. For water-splitting experiments, steam was fed to the inner side of the
membrane at a rate of 140 mL min−1 (5 g h−1 water to steam generator), carried in 80-100 mL
min−1 Ar. For co-feed experiments, the same flow of steam was carried by 75-200 mL min-1 CO2.
At the outer side of the membrane, the reactor shell was purged with 200-1000 mL min−1 Ar. The
compositions of both gas streams exiting the solar reactor were analyzed simultaneously using two
gas chromatographs (GC). Steady state was defined as the condition at which the measured gas
concentration was within 2% of the mean over the previous five consecutive measurements
collected at a frequency of one every two minutes:
(5)
 
 
 
1
5
1
5
1
50.02
1
5
n
i n i j
j n
n
i j
j n
c t c t
c t
 
 
where ci(tj) is the volumetric concentration of species i at time point j. Steady-state data was
collected for at least 19 minutes and the arithmetic mean was used to summarize the results at
each experimental condition.
Thermodynamic Analysis
We calculated the thermodynamic equilibrium limits of thermolysis of CO2 and H2O in a
membrane reactor to compare to experimental results. Relatively fast rates are expected for each
reaction step: gas-phase thermolysis, heterogeneous surface reaction, and oxygen bulk diffusion.
When each of these serial processes are sufficiently fast, the global kinetics are fast, and
thermodynamics govern the net reaction. In this case, kinetics can be neglected.
Page 8 of 21Reaction Chemistry & Engineering
Reaction Chemistry & Engineering Accepted Manuscript
Open Access Article. Published on 14 March 2019. Downloaded on 4/14/2019 7:04:04 PM.
This article is licensed under a
Creative Commons Attribution 3.0 Unported Licence.
View Article Online
DOI: 10.1039/C8RE00218E
9
Previous observations with a solar cavity-receiver containing a porous ceria structure directly
exposed to high-flux irradiation reported that the overall kinetics are not controlled by solid-state
diffusion within the crystal lattice.33 This is also expected for a ceria membrane because the
measured values of ambipolar diffusion coefficients of oxygen in ceria (1.5×10-5-4×10-4 cm2 s-1
in the range 1673-1823 K22) translate to diffusion times on the order of seconds for the length
scales across the 0.5 mm-thick membrane. Thus, as far as solid-state diffusion is concerned, the
transport of oxygen vacancies through the membrane is almost instantaneous compared with the
time scales of data collection.
Reaction rates have an exponential dependence on temperature, scaling with exp(-EA/RT), as
seen in the Arrhenius equation. The high temperature in the range 1723-1873 K and consequent
high activity of reactive oxygen vacancies at the surface of the membrane are expected to lead to
fast surface exchange of oxygen from the gas into the solid phase.19 While studies of non-
isothermal processes show that heating rate limits the oxygen release rate, heat transfer should
not be limiting in this isothermal process because the heat of reaction is much lower than the heat
input.23, 33 Likewise, the high temperature, along with the catalytic effect of ceria, implies very
fast thermolysis reactions in the gas-phase.34 Therefore, a purely thermodynamic model is
expected to adequately predict reactor performance.
However, to reach the thermodynamic limit in the countercurrent-flow reactor, there must be
sufficient membrane area and sweep gas relative to the flowrate of reactant. To account for the
oxygen capacity in a given flow of sweep gas, a thermodynamic model described by Bulfin was
applied, which is specific to countercurrent-flow reactors.35 This approach guarantees
compliance with the second law of thermodynamics and conservation of mass along the entire
reactor by means of a dimensionless oxygen exchange coordinate, κ, defined as the number of
moles of O2 crossing the membrane up to a certain point along the length, x, per mole of oxidant
fed:
(6)
 
2
O
0
oxidant
( ) ,
xj x dx
xn
&
where oxidant is the molar flow rate of H2O and/or CO2 (CO2+H2O), and jO2(x) is the molar flux of
O2 from the oxidant flow to the sweep gas as a function of the length along the membrane. Then
Page 9 of 21 Reaction Chemistry & Engineering
Reaction Chemistry & Engineering Accepted Manuscript
Open Access Article. Published on 14 March 2019. Downloaded on 4/14/2019 7:04:04 PM.
This article is licensed under a
Creative Commons Attribution 3.0 Unported Licence.
View Article Online
DOI: 10.1039/C8RE00218E
10
pO2 in each flow can be formulated as a function of κ, and pO2,oxidant is determined by the
thermodynamic equilibrium of thermolysis, as described by equations 3 and 4 in the
introduction. For O2 to spontaneously flow from the oxidant flow to the sweep gas, partial
pressures must satisfy pO2,oxidant(κ) > pO2,sweep(κ), and equilibrium is reached if there exists a value
of κ in [0, κtotal] where pO2,oxidant(κ) = pO2,sweep(κ). Bulfin details the full methodology in an
example tailored to thermolysis in a membrane reactor with a counter-current sweep flow.35
The countercurrent-flow thermodynamic model was implemented and solved numerically in
Matlab using thermodynamic data from NIST JANAF.36 The model input parameters are T, p,
pO2, the relative flow of sweep to oxidant (sweep/oxidant), and the relative flow rates of CO2 and
H2O (CO2/H2O). Note that pO2 refers to the O2 impurity at the inlet of the sweep gas. The values
were generally chosen to match experimental conditions as determined from mass flow
controller, thermocouple, and GC measurements. For example, the reactor operated at ambient
pressure and 1 bar total pressure was set on both sides of the membrane. It was not possible to
measure the reaction temperature of the gas inside the membrane. Instead, the measured
temperature at the outer wall of the shell tube (certainly greater than the reaction temperature)
and a simple heat transfer model were used to estimate the reaction temperature. Consequently,
this analysis presents the thermodynamic limits as a shaded region, where the upper and lower
bounds are the limits at the maximum measured shell temperature and the estimated reaction
temperature, respectively. In the range of operation, the difference between these temperatures is
approximately 50 K. The simple heat transfer model used to calculate this temperature difference
is described in the Electronic Supplemental Information (ESI). The outputs of the
thermodynamic model are the partial pressures of the products and conversion of the reactants at
equilibrium.
Results and Discussion
Figure 3 summarizes the steady-state specific production rates measured experimentally for fuel,
comprising of CO (light shading) and H2 (dark shading), and O2 as a function of three process
variables: (a) time, (b) T, and (c) pO2. These experiments used a roughly equimolar feed of CO2
and H2O and produced fuel with a relatively higher fraction of CO than H2. This product
proportion is consistent with the more favorable change in Gibbs free energy for thermolysis of
Page 10 of 21Reaction Chemistry & Engineering
Reaction Chemistry & Engineering Accepted Manuscript
Open Access Article. Published on 14 March 2019. Downloaded on 4/14/2019 7:04:04 PM.
This article is licensed under a
Creative Commons Attribution 3.0 Unported Licence.
View Article Online
DOI: 10.1039/C8RE00218E
11
CO2 at high temperatures seen in Figure 1a. The average molar ratio O2:fuel over 19 experiments
was 0.53 ± 0.07, corroborating a closed mass balance. Furthermore, no other byproducts were
detected, indicating the absence of undesired side reactions. Specifically, Figure 3a shows
production rates over time at steady-state at a nominal T of 1873 K and pO2 of 0.4 Pa. Note that T
refers to the maximum measured shell temperature and pO2 refers to the partial pressure of O2 at
the Ar inlet, which is the minimum pO2 in the system and determined by the impurity in the
sweep gas. Gas evolution rates leveled off at a constant incident radiative flux of about 3500
suns (1 sun = 1 kW/m2) over the cavity’s aperture, and thus the reactor demonstrated continuous,
steady-state operation at isothermal conditions. Figure 3b shows the steady-state average
production rates as a function of T in the range 1723-1873 K. Mass flow rates were kept constant
at 5 g h-1 H2O with 100 mL min-1 CO2 to the inner side and 500 mL min-1 Ar to the outer side of
the membrane (L denotes standard liters). The measured pO2 ranged from 0.4 – 1.7 Pa due to
variation in the small amount of air leakage into the reactor during different experimental runs.
The specific fuel and O2 production rates at steady state increased with temperature, in
accordance with the thermodynamic dependence on exp(-ΔG°/RT). Finally, Figure 3c shows the
steady-state average production rates as a function of pO2 in the range 0.2 – 0.9 Pa at 1873 K.
Although the range of pO2 tested was small, the gas production rates indeed decreased at higher
pO2, as expected from the equilibrium relationship for the thermolysis of CO2 and H2O described
in equations 3 and 4. However, to achieve this range of pO2, the flow rate of Ar was adjusted,
thus changing the relative flow of sweep gas, sweep/oxidant, which is also a key thermodynamic
parameter in this reactor system.
A B C
Page 11 of 21 Reaction Chemistry & Engineering
Reaction Chemistry & Engineering Accepted Manuscript
Open Access Article. Published on 14 March 2019. Downloaded on 4/14/2019 7:04:04 PM.
This article is licensed under a
Creative Commons Attribution 3.0 Unported Licence.
View Article Online
DOI: 10.1039/C8RE00218E
12
Fig. 3 Trends in steady-state specific gas production rates. (a) Steady-state production rates of
fuel (CO and H2) and O2 vs. time at 1873 K and 0.4 Pa O2. Steady-state average production rates
of fuel and O2 vs. (b) T, keeping all mass flow rates constant (c) pO2, keeping T constant at 1873
K. Feed was equimolar CO2 and H2O in all cases. Fuel production is composed of two
contributions distinguished by shade: CO (light) and H2 (dark).
Figure 4 shows the experimental steady-state conversion of CO2 to CO and H2O to H2 as a
function of (a) T in the range 1723-1873 K at pO2 = 1 Pa, sweep/oxidant = 2.4, and (b) sweep/oxidant
ranging 1-5 at T = 1873 K, pO2 = 0.5 Pa. The thermodynamic limits are also indicated, denoted by
a band whose upper and lower bounds are calculated as described above at Tshell and the
estimated reaction temperature, Tshell – 50 K, respectively. The band is wider for CO2 than H2O
because its reaction favorability changes more steeply with T. Like the gas production rates,
conversion of reactants in Figure 4a increased with T, as expected. The conversion also increased
at higher relative sweep rates in Figure 4b, because sweep/oxidant determines the total amount of
O2 that can be removed across the membrane at equilibrium. In fact, the thermodynamic analysis
reveals that the trend in production rates observed in Figure 3c is more attributable to varying
sweep/oxidant than pO2. The pO2 in the sweep gas must be lower than pO2 in the oxidant stream at all
points along the membrane to drive transfer of O2. At sufficiently low pO2, however, this
parameter does not have a strong influence on the conversion of reactants, shown in Figure S3 in
the ESI. In the range of pO2 observed in experiments (0.2-0.9 Pa), the theoretical conversion of
each reactant at constant sweep/oxidant is almost flat, while the experimental data points exhibit a
trend due to varying relative sweep rates. In contrast, the same experimental data plotted against
sweep/oxidant in Figure 4b match the shape of the equilibrium limit curves.
In general, the results indicate that the reactor performance indeed approaches the
thermodynamic limit for a countercurrent flow reactor. Importantly, the experimental conversion
does not exceed the theoretical limit. Furthermore, the experimental points lie closer to the lower
bound of the equilibrium region, suggesting that the simple heat transfer model is necessary and
effective to estimate the reaction temperature. While the experimental conversion of H2O closely
follows the lower bound of the predicted thermodynamic limit, the experimental conversion of
CO2 falls short. The discrepancy is less than a factor of two and may be a result of the water-gas
Page 12 of 21Reaction Chemistry & Engineering
Reaction Chemistry & Engineering Accepted Manuscript
Open Access Article. Published on 14 March 2019. Downloaded on 4/14/2019 7:04:04 PM.
This article is licensed under a
Creative Commons Attribution 3.0 Unported Licence.
View Article Online
DOI: 10.1039/C8RE00218E
13
shift (WGS) reaction occurring at lower T downstream of the reactor, before the GC analysis.
WGS consumes some CO to produce additional H2 and is thus the difference of the CO2 and
H2O dissociation reactions (Eq. 2 minus Eq. 1):
(7)
2 2 2
CO + H O H + CO
In this case, the GC measurements may not be representative of the composition in the reactor;
the actual conversion of CO2 may have been higher, and the conversion of H2O lower,
potentially equalizing the difference seen in Figure 4 between the experimental data and the limit
for each gas.
Page 13 of 21 Reaction Chemistry & Engineering
Reaction Chemistry & Engineering Accepted Manuscript
Open Access Article. Published on 14 March 2019. Downloaded on 4/14/2019 7:04:04 PM.
This article is licensed under a
Creative Commons Attribution 3.0 Unported Licence.
View Article Online
DOI: 10.1039/C8RE00218E
14
Fig. 4 Comparison of average experimental conversion of CO2 and H2O at steady state versus
equilibrium limits as a function of (a) T, at sweep/oxidant = 2.4 and 1 Pa O2, (b) sweep/oxidant, at
1873 K and 0.5 Pa O2.
Figure 5 illustrates the effect of the molar feed ratio, CO2:H2O, on the: (a) steady-state average
production rates of fuel and O2, (b) conversions of CO2 and H2O, and (c) molar ratio of the two
fuels produced, CO:H2. Figures 4b and c compare the results measured experimentally with
those predicted from thermodynamics. The relative flow rates of CO2 and H2O were varied while
maintaining steam feed rate at 5 g h-1 H2O, temperature at 1873 K, and pO2 at 0.5 Pa. Figure 4a
shows that the overall fuel production (sum of H2 and CO) increased with CO2:H2O, as did the
production rate of CO, which occurred because the total feed rate and CO2 feed rate both
increased. The production rate of H2, on the other hand, decreased with CO2:H2O. It must be
emphasized that sweep/oxidant did not remain constant over experimentation, but rather decreased
with CO2:H2O because the flow rate of sweep gas remained constant. A constant flowrate of
sweep gas is less effective at maintaining low pO2 as the amount of O2 to be removed increases.
In the case of CO, the effect of increasing CO2 feed rate compensated for the decreasing sweep
ratio, so that the net production rate increased. However, the H2 production decreased.
The confounding factors of changing both the total feed rate and the relative sweep rate are
accounted for by plotting conversion instead of production rate in Figure 5b. As in Figure 4, the
absolute values of H2O conversion match the model results better than for CO2. The
experimental conversions of both CO2 and H2O decreased slightly with CO2:H2O, in agreement
with the trend predicted at equilibrium. There are two contributions to the negative trend in
conversion. First, as already mentioned, sweep/oxidant decreased with CO2:H2O, which decreased
the conversion. However, conversion of each reactant is predicted to decrease slightly with
CO2:H2O even with a constant sweep/oxidant. The second, smaller reason for the trend is related to
the difference in favorability of thermolysis of CO2 and H2O. As the feed ratio increases, a
higher proportion of the feed is CO2, which has a higher conversion than H2O at these
conditions. In fact, the overall conversion of reactants to products actually increased slightly with
CO2:H2O. However, the O2 capacity of the sweep gas was unchanged and therefore the
conversion of each individual reactant must decrease to balance the production of O2 with its
removal. This result indicates that a feed of CO2 requires a higher relative sweep rate than an
Page 14 of 21Reaction Chemistry & Engineering
Reaction Chemistry & Engineering Accepted Manuscript
Open Access Article. Published on 14 March 2019. Downloaded on 4/14/2019 7:04:04 PM.
This article is licensed under a
Creative Commons Attribution 3.0 Unported Licence.
View Article Online
DOI: 10.1039/C8RE00218E
15
equal feed of H2O, because its higher potential conversion leads to a larger amount of O2 that
must be removed.
Consistent with Figure 5a, Figure 5c shows that the product ratio CO:H2 increased with feed
ratio CO2:H2O, as expected intuitively. Interestingly, CO:H2 is always greater than the
corresponding CO2:H2O, which further confirms that dissociation of CO2 is more favorable than
that of H2O at equivalent conditions. The observed experimental trend qualitatively matches
equilibrium thermodynamics, though with a smaller slope because the experimental conversion
of CO2 is lower than predicted. In consideration of downstream processing, the Fischer-Tropsch
synthesis favors a syngas fed with 1:2 moles CO:H2.38, 39 According to Figure 5c, the product
ratio can be adjusted via the feed ratio, and a 1:2 product ratio would require a feed ratio smaller
than the minimum tested here.
Page 15 of 21 Reaction Chemistry & Engineering
Reaction Chemistry & Engineering Accepted Manuscript
Open Access Article. Published on 14 March 2019. Downloaded on 4/14/2019 7:04:04 PM.
This article is licensed under a
Creative Commons Attribution 3.0 Unported Licence.
View Article Online
DOI: 10.1039/C8RE00218E
16
Fig. 5 Effect of molar feed ratio CO2:H2O on: (a) steady-state average production rate of fuel and
O2, where fuel is composed of two contributions distinguished by shade: CO (light) and H2
(dark); (b) molar conversion of CO2 and H2O calculated from experiments and their limits at
equilibrium; and (c) molar fuel ratio CO:H2 measured experimentally and at equilibrium.
Reaction conditions were sweep/oxidant varying from 1.6-2.5 and constant 5 g h-1 H2O feed rate,
1873 K, and 0.5 Pa O2.
Page 16 of 21Reaction Chemistry & Engineering
Reaction Chemistry & Engineering Accepted Manuscript
Open Access Article. Published on 14 March 2019. Downloaded on 4/14/2019 7:04:04 PM.
This article is licensed under a
Creative Commons Attribution 3.0 Unported Licence.
View Article Online
DOI: 10.1039/C8RE00218E
17
The maximum conversions observed experimentally were 1.0% CO2 and 0.4% H2O at 1873 K,
sweep/oxidant = 5, and 0.2 Pa O2 (0.7% overall conversion of reactants). In general, the absolute
values of both theoretical and experimental conversion were lower in the co-feed case tested here
than in the pure-CO2 feed case tested previously,12 because the relative sweep rates were lower in
this set of experiments. The base case sweep/oxidant was 8 in pure-CO2 experiments and 2.4 in
these co-feed experiments. The relative sweep rate is a significant thermodynamic parameter for
sweep gas operation, and the application of the countercurrent flow model was essential to
accurately predict the behavior of the reactor. Furthermore, although there was no effort to
optimize the efficiency of the reactor in these proof-of-concept experiments, sweep/oxidant also
impacts efficiency because it determines how much sweep gas must be heated and circulated per
unit fuel produced.
Conclusion
We have demonstrated steady-state splitting of a mixed feed containing CO2 and H2O into
separate streams of syngas fuel and O2 using an isothermal tubular ceria membrane reactor
driven by simulated concentrated solar radiation. The experimental results generally agreed with
trends predicted by thermodynamics. The conversion of CO2 to CO was favored over H2O to H2,
consistent with the energetics of the respective thermolysis reactions. The co-thermolysis of a
mixture of CO2 and H2O is more complex than feeding either CO2 or H2O separately to the
reactor, both experimentally and in the theoretical analysis. In the co-feed case, the mixture of
CO2, H2O, CO, and H2 could undergo additional reactions, especially the reverse water-gas shift
(RWGS). As seen in equation 7, the RWGS reaction is not independent of the two thermolysis
reactions, and thus does not change the equilibrium conversion of each reactant from the pure-
feed values. Other possible reactions were found to be negligible from the lack of byproducts
predicted by thermodynamics and measured experimentally at these conditions. Absence of
byproducts and a 2:1 ratio of fuel:oxygen together confirmed 100% selectivity for the desired
splitting reactions.
The solar thermochemical membrane reactor unifies both CO2 and H2O splitting in a single
modular and scalable device and offers a technically viable pathway to single-step syngas
production. However, determining an appropriate relative sweep rate is challenging in co-feed
Page 17 of 21 Reaction Chemistry & Engineering
Reaction Chemistry & Engineering Accepted Manuscript
Open Access Article. Published on 14 March 2019. Downloaded on 4/14/2019 7:04:04 PM.
This article is licensed under a
Creative Commons Attribution 3.0 Unported Licence.
View Article Online
DOI: 10.1039/C8RE00218E
18
operation because the different favorability for thermolysis of CO2 and H2O implies different
optimums for each species. In addition, these energetic differences mean that H2O needs to be
fed in large excess to achieve a desirable syngas composition. Therefore, it may still be attractive
to produce CO and H2 separately and mix them into syngas as needed. Furthermore, the single-
step approach embodied in the membrane reactor must compete with multistep cycles currently
available. Thus, further R&D and alternative membrane configurations are needed to boost mass
conversions and consequently reach favorable solar-to-fuel energy efficiencies, a challenge
because T and pO2 determine the thermodynamic limits.
Conflicts of Interest
There are no conflicts to declare.
Acknowledgements
This work was funded by the Swiss National Science Foundation (Ambizione Energy Grant No.
166883) and by the Chinese Academy of Sciences (International Collaboration Key Program
award no. 182211KYSB20160043). We thank Brendan Bulfin for fruitful discussions and
Patrick Basler and Simon Minder for supporting the experimental campaign.
References
1. N. S. Lewis, Science, 2016, 351, aad1920.
2. G. A. Ozin, Advanced Materials, 2015, 27, 1957-1963.
3. J. A. Herron, J. Kim, A. A. Upadhye, G. W. Huber and C. T. Maravelias, Energy
Environ. Sci., 2015, 8, 126-157.
4. D. Marxer, P. Furler, J. Scheffe, H. Geerlings, C. Falter, V. Batteiger, A. Sizmann and A.
Steinfeld, Energy & Fuels, 2015, 29, 3241-3250.
5. W. C. Chueh, C. Falter, M. Abbott, D. Scipio, P. Furler, S. M. Haile and A. Steinfeld,
Science, 2010, 330, 1797-1801.
6. M. Romero and A. Steinfeld, Energy & Environmental Science, 2012, 5, 9234.
7. Y. Hao, C. K. Yang and S. M. Haile, Phys Chem Chem Phys, 2013, 15, 17084-17092.
8. C. L. Muhich, B. W. Evanko, K. C. Weston, P. Lichty, X. Liang, J. Martinek, C. B.
Musgrave and A. W. Weimer, Science, 2013, 341, 540-542.
9. L. J. Venstrom, R. M. De Smith, Y. Hao, S. M. Haile and J. H. Davidson, Energy &
Fuels, 2014, 28, 2732-2742.
10. E. A. Fletcher and R. L. Moen, Science, 1977, 197, 1050-1056.
11. J. E. Noring, R. B. Diver and E. A. Fletcher, Energy, 1981, 6, 109-121.
Page 18 of 21Reaction Chemistry & Engineering
Reaction Chemistry & Engineering Accepted Manuscript
Open Access Article. Published on 14 March 2019. Downloaded on 4/14/2019 7:04:04 PM.
This article is licensed under a
Creative Commons Attribution 3.0 Unported Licence.
View Article Online
DOI: 10.1039/C8RE00218E
19
12. M. Tou, R. Michalsky and A. Steinfeld, Joule, 2017, 1, 146-154.
13. A. Evdou, L. Nalbandian and V. Zaspalis, Journal of Membrane Science, 2008, 325, 704-
711.
14. W. Jin, C. Zhang, X. Chang, Y. Fan, W. Xing and N. Xu, Environmental Science &
Technology, 2008, 42, 3064-3068.
15. L. Nalbandian, A. Evdou and V. Zaspalis, International Journal of Hydrogen Energy,
2009, 34, 7162-7172.
16. H. Wang, Y. Hao and H. Kong, International Journal of Energy Research, 2015, 39,
1790-1799.
17. X. Y. Wu, L. Chang, M. Uddi, P. Kirchen and A. F. Ghoniem, Phys. Chem. Chem. Phys.,
2015, 17, 10093-10107.
18. L. Zhu, Y. Lu and S. Shen, Energy, 2016, 104, 53-63.
19. T. C. Davenport, C.-K. Yang, C. J. Kucharczyk, M. J. Ignatowich and S. M. Haile,
Energy Technology, 2016, 4, 764-770.
20. P. Furler, J. R. Scheffe and A. Steinfeld, Energy & Environmental Science, 2012, 5,
6098-6103.
21. P. M. Geffroy, J. Fouletier, N. Richet and T. Chartier, Chemical Engineering Science,
2013, 87, 408-433.
22. S. Ackermann, J. R. Scheffe and A. Steinfeld, The Journal of Physical Chemistry C,
2014, 118, 5216-5225.
23. W. C. Chueh and S. M. Haile, Philosophical Transactions of the Royal Society A, 2010,
368, 3269-3294.
24. M. Mogensen, N. M. Sammes and G. A. Tompsett, Solid State Ionics, 2000, 129, 63-94.
25. A. Glück, R. Tamme, H. Kalfa and C. Streuber, Solar Energy Materials, 1991, 24, 240-
248.
26. R. Bader, L. J. Venstrom, J. H. Davidson and W. Lipiński, Energy & Fuels, 2013, 27,
5533-5544.
27. S. Brendelberger, H. von Storch, B. Bulfin and C. Sattler, Solar Energy, 2017, 141, 91-
102.
28. M. Ezbiri, K. M. Allen, M. E. Galvez, R. Michalsky and A. Steinfeld, ChemSusChem,
2015, 8, 1966-1971.
29. B. Bulfin, J. Vieten, C. Agrafiotis, M. Roeb and C. Sattler, Journal of Materials
Chemistry A, 2017, 5, 18951-18966.
30. W. He, H. Huang, J.-f. Gao, L. Winnubst and C.-s. Chen, Journal of Membrane Science,
2014, 452, 294-299.
31. X. Tan, Y. Liu and K. Li, Ind. Eng. Chem. Res., 2005, 44, 61-66.
32. W. T. Welford and R. Winston, High Collection Nonimaging Optics, Academic Press,
Inc., San Diego, 1989.
33. P. Furler, J. Scheffe, M. Gorbar, L. Moes, U. Vogt and A. Steinfeld, Energy & Fuels,
2012, 26, 7051-7059.
34. Q. Jiang, Z. Chen, J. Tong, M. Yang, Z. Jiang and C. Li, Chem Commun (Camb), 2017,
53, 1188-1191.
35. B. Bulfin, Physical Chemistry Chemical Physics, 2019, 21, 2186-2195.
36. M. W. Chase, C.A. Davies, J. R. Downey, D.J. Frurip, R. A. McDonald and A. N.
Syverud, Journal, 1998, DOI: 10.18434/T42S31.
37. A. Roine, HSC Chemistry 5, Outokumpu Research Oy, Pori, Finland, 2002.
Page 19 of 21 Reaction Chemistry & Engineering
Reaction Chemistry & Engineering Accepted Manuscript
Open Access Article. Published on 14 March 2019. Downloaded on 4/14/2019 7:04:04 PM.
This article is licensed under a
Creative Commons Attribution 3.0 Unported Licence.
View Article Online
DOI: 10.1039/C8RE00218E
20
38. H. Schulz, Applied Catalysis A: General, 1999, 186, 3-12.
39. D. Leckel, Energy & Fuels, 2009, 23, 2342-2358.
Page 20 of 21Reaction Chemistry & Engineering
Reaction Chemistry & Engineering Accepted Manuscript
Open Access Article. Published on 14 March 2019. Downloaded on 4/14/2019 7:04:04 PM.
This article is licensed under a
Creative Commons Attribution 3.0 Unported Licence.
View Article Online
DOI: 10.1039/C8RE00218E
Tou et al
Table of Contents entry
First-time experimental demonstration of simultaneous thermolysis of CO2 and H2O in a solar-
driven membrane reactor
Page 21 of 21 Reaction Chemistry & Engineering
Reaction Chemistry & Engineering Accepted Manuscript
Open Access Article. Published on 14 March 2019. Downloaded on 4/14/2019 7:04:04 PM.
This article is licensed under a
Creative Commons Attribution 3.0 Unported Licence.
View Article Online
DOI: 10.1039/C8RE00218E
... A cerium-based thermochemical reactor was integrated with a concentrated solar collector, characterized by 4200 suns. This resulted in the development of a single, modular, and scalable device, and a suitable experimental setup was constructed to test the system under steady-state conditions [30]. The use of indirect solar irradiation and ceria particles in a fluidized bed reactor enhances the hydrogen yield and reduces the hydrogen price [31]. ...
Article
This paper presents a novel dynamic simulation model for assessing the energy performance of solar-driven systems employed in green hydrogen production. The system consists of a parabolic dish collector that focuses solar radiation on two cerium-based thermochemical reactors. The model is based on a transient finite-difference method to simulate the thermal behaviour of the system and it integrates a theoretical analysis of materials and operating principles. Different empirical data were considered for experimentally validating it: a good agreement between experimental and simulated results was obtained for the temperatures calculated inside the thermochemical reactor (R2 = 0.99, MAPE = 6.3%) and the hourly flow rates of hydrogen, oxygen, and carbon monoxide (R2 = 0.96, MAPE = 10%) inside the thermochemical reactor. The model was implemented in a MatLab tool for the system dynamic analysis under different boundary conditions. Subsequently, to explore the capability of this approach, the developed tool was used for analysing the examined device operating in twelve different weather zones. The obtained results comprise heat maps of specific crucial instants and hourly dynamic trends showing redox reaction cycles occurring into the thermochemical reactors. The yearly hydrogen production ranges from 1.19 m3/y to 1.64 m3/y according to the hourly incident solar radiations, outdoor air temperatures and wind speeds. New graphic tools for rapid feasibility studies are presented. The developed tools and the obtained results can be useful to the basic design of this technology and for the multi-objective optimization of its layout and main design/operating parameters.
... Tou et al. also tested the feasibility of splitting both H 2 O and CO 2 in a tubular CeO 2 membrane reactor heated by simulated concentrated solar radiation into separate streams of oxygen and syngas, respectively. 77 This simple solar membrane reactor technology demonstrates its potential for co-splitting CO 2 and H 2 O to fuels under high flux conditions. However, heat and mass transfer were impractically slow, restricting the reaction rates. ...
Article
Full-text available
The integration of membrane separation processes with chemical reactions through oxygen transport dual‐functional membrane reactors has attracted significant attention due to the potential for process intensification, which also can create a synergy between the two units. This approach holds promise for promoting green chemistry principles by reducing energy consumption and environmental pollution. Despite its potential, a comprehensive review of recent advancements exploring the full potential of oxygen transport dual‐functional membrane reactors (coupling two distinct reactions) in enhancing membrane performance is currently lacking. To address this gap, this perspective article presents various concepts and principles of oxygen transport dual‐functional membrane reactors and provides an overview of recent advances and applications. Additionally, the challenges and opportunities for future research to enhance the efficiency of the process toward industrialization are discussed and highlighted. These include developing novel oxygen transport membrane materials, optimizing membrane engineering, innovating membrane reactor design, and exploring new applications and reaction mechanisms.
... Tou et al. [67] developed a solar-driven membrane reactor. They tested it for syngas production via simultaneous WS and CDS. ...
Article
Full-text available
Metal oxide (MO) based solar thermochemical H2O (WS) and CO2 splitting (CDS) is one of the most promising and potential-containing processes that can be used to produce H2 and syngas (liquid fuel precursor). Several non-volatile and volatile MOs were considered redox materials for the solar-driven WS and CDS operation. Among all the examined redox materials, based on their high O2 storage capacity, faster oxidation kinetics, and good stability, ceria and doped ceria materials are deemed to be one of the best alternatives for the operation of the thermochemical redox reactions associated with the WS and CDS. Pure ceria was used for solar fuel production for the first time in 2006. A review paper highlighting the work done on the ceria-based solar thermochemical redox WS and CDS cycle from 2006 until 2016 is already published elsewhere by the author. This review paper presents all the significant findings reported in applying pure ceria and doped ceria materials for the WS and CDS by research teams worldwide.
... This compound has the greatest reaction rate, maximum O 2 exchange capacity, and minimum reaction temperature. Tou et al. demonstrated the first solar-based CO 2 and H 2 O co-thermolysis (Tou et al., 2019). Gaseous H 2 O/CO 2 mixtures were passed across a non-stoichiometric ceria (CeO 2−δ ) membrane with 4200 suns of solar focused radiation. ...
... Ceria membranes coated with perovskite were investigated on-sun under various operating conditions [158], showing improved CO 2 dissociation rates. Membranes coated with thin perovskite layers featuring enhanced ionic conductivity and oxygen partial pressure gradient enable isothermal CO 2 splitting (1450-1550 • C) with high dissociation rates in comparison with pristine ceria membranes [159,160]. On one side of the membrane (feed side), flowing CO 2 was dissociated and on the other side (sweep side), the oxygen release was favored through an inert gas sweeping. Mixed ionic-electronic conducting (MIEC) membranes for the thermochemical reduction of CO 2 have been developed and are further reviewed in [161]. ...
Article
Full-text available
Redox materials have been investigated for various thermochemical processing applications including solar fuel production (hydrogen, syngas), ammonia synthesis, thermochemical energy storage, and air separation/oxygen pumping, while involving concentrated solar energy as the high-temperature process heat source for solid–gas reactions. Accordingly, these materials can be processed in two-step redox cycles for thermochemical fuel production from H2O and CO2 splitting. In such cycles, the metal oxide is first thermally reduced when heated under concentrated solar energy. Then, the reduced material is re-oxidized with either H2O or CO2 to produce H2 or CO. The mixture forms syngas that can be used for the synthesis of various hydrocarbon fuels. An alternative process involves redox systems of metal oxides/nitrides for ammonia synthesis from N2 and H2O based on chemical looping cycles. A metal nitride reacts with steam to form ammonia and the corresponding metal oxide. The latter is then recycled in a nitridation reaction with N2 and a reducer. In another process, redox systems can be processed in reversible endothermal/exothermal reactions for solar thermochemical energy storage at high temperature. The reduction corresponds to the heat charge while the reverse oxidation with air leads to the heat discharge for supplying process heat to a downstream process. Similar reversible redox reactions can finally be used for oxygen separation from air, which results in separate flows of O2 and N2 that can be both valorized, or thermochemical oxygen pumping to absorb residual oxygen. This review deals with the different redox materials involving stoichiometric or non-stoichiometric materials applied to solar fuel production (H2, syngas, ammonia), thermochemical energy storage, and thermochemical air separation or gas purification. The most relevant chemical looping reactions and the best performing materials acting as the oxygen carriers are identified and described, as well as the chemical reactors suitable for solar energy absorption, conversion, and storage.
... The obtained performances were similar to those of ceria foams (~200-300 µmol/g of H2/CO, peak production rate: 9.5 mL/min/g) under comparable cycling conditions. Alternatively, a membrane solar reactor was developed for continuous and isothermal splitting of CO2 (and H2O) under a gradient of O2 partial pressure across the membrane [138,[201][202][203][204]. As described in Figure 8, the original process was demonstrated in a solar reactor under high-flux solar irradiation using a dense ceramic membrane with tubular shape (mixed ionic-electronic conducting materials, MIEC) that separates a reduction zone at low pO2 (sweep side) and an oxidation zone fed with reactive gas flow (feed side). ...
Article
Full-text available
The solar thermochemical two-step splitting of H2O and CO2 based on metal oxide compounds is a promising path for clean and efficient generation of hydrogen and renewable synthetic fuels. The two-step process is based on the endothermic solar thermal reduction of a metal oxide releasing O2 using a high-temperature concentrated solar heat source, followed by the exothermic oxidation of the reduced oxide with H2O and/or CO2 to generate pure H2 and/or CO. This pathway relates to one of the emerging and most promising processes for solar thermochemical fuel production encompassing green H2 and the recycling/valorization of anthropogenic greenhouse gas emissions. It represents an efficient route for solar energy conversion and storage into renewable and dispatchable fuels, by directly converting the whole solar spectrum using heat delivered by concentrating systems. This eliminates the need for photocatalysts or intermediate electricity production, thus bypassing the main limitations of the low-efficient photochemical and electrochemical routes currently seen as the main green methods for solar fuel production. In this context, among the relevant potential redox materials, thermochemical cycles based on volatile and non-volatile metal oxides are particularly attractive. Most redox pairs in two-step cycles proceed with a phase change (solid-to-gas or solid-to-liquid) during the reduction step, which can be avoided by using non-stoichiometric oxides (chiefly, spinel, fluorite, or perovskite-structured materials) through the creation of oxygen vacancies in the lattice. The oxygen sub-stoichiometry determines the oxygen exchange capacity, thus determining the fuel production output per mass of redox-active material. This paper provides an overview of the most advanced cycles involving ZnO/Zn, SnO2/SnO, Fe3O4/FeO, ferrites, ceria, and perovskites redox systems by focusing on their ability to perform H2O and CO2 splitting during two-step thermochemical cycles with high fuel production yields, rapid reaction rates, and performance stability. Furthermore, the possible routes for redox-active material integration and processing in various solar reactor technologies are also described.
Article
Full-text available
Catalytic activity is affected by surface morphology, and specific surfaces display greater activity than others. A key challenge is to define synthetic strategies to enhance the expression of more active...
Article
Full-text available
Solar-driven thermochemical conversion of H2O and CO2 into sustainable fuels, based on redox cycle, provides a promising path for alternative energy, as it employs the solar energy as high-temperature heat supply and adopts H2O and CO2 as initial feedstock. This review describes the sustainable fuels production system, including a series of physical and chemical processes for converting solar energy into chemical energy in the form of sustainable fuels. Detailed working principles, redox materials, and key devices are reviewed and discussed to provide systematic and in-depth understanding of thermochemical fuels production with the aid of concentrated solar power technology. In addition, limiting factors affecting the solar-to-fuel efficiency are analyzed; meanwhile, the improvement technologies (heat recovery concepts and designs) are summarized. This study therefore sets a pathway for future research works based on the current status and demand for further development of such technologies on a commercial scale.
Article
Full-text available
Hydrogen (H2) produced from renewables will have a growing impact on the global energy dynamics towards sustainable and carbon‐neutral standards. The share of green H2 is still too low to meet the net‐zero target, while the demand for high‐quality hydrogen continues to rise. These factors amplify the need for economically viable H2 generation technologies. The present article aims at evaluating the existing technologies for high‐quality H2 production based on solar energy. Technologies such as water electrolysis, photoelectrochemical and solar thermochemical water splitting, liquid metal reactors and plasma conversion utilize solar power directly or indirectly (as carbon‐neutral electrons) and are reviewed from the perspective of their current development level, technical limitations and future potential.
Article
Full-text available
Countercurrent reactors can be utilized in chemical reaction systems which involve either a reaction between flows of different phases, or reactions between flows separated by a selective permeable membrane. This idea is quite similar in nature to a countercurrent heat exchanger, where the inlet of one participating flow is exposed to the outlet of the opposite flow. A countercurrent configuration can therefore improve the reaction conversion extent and transport properties. Here we formulate a straightforward approach in terms of an exchange coordinate, in order to determine an upper bound of species exchange in such systems, subject to the second law of thermodynamics and conservation of mass. The methodology is independent of the specifics of reactor design and can be generally applied to determine the maximum thermodynamic benefit of using a countercurrent reactor. We then demonstrate the analysis for a number of thermochemical fuel production routes; membrane thermolysis of carbon dioxide, dry methane reforming across a membrane, reverse water gas shift across a membrane, and the thermochemical ceria cycle.
Article
Full-text available
Two step metal oxide thermochemical redox cycles have seen growing interest in the research community over the last two decades. In particular, they have often been studied as a means of converting heat from concentrated solar power to chemical energy, which can subsequently be used for a number of applications. In this work, we offer critical perspective and valuable insight on these research fields, from authors with a combined experience of 50+ years in metal oxide redox cycles. The current fundamental understanding of thermochemical redox materials, and the implications and limitations this has on the redox thermodynamics are discussed. The underlying fundamentals of the redox materials are then used to give insight into the theoretical limitations imposed on a number of applications including; solar thermochemical fuel production, solar energy storage for off sun power generation, thermochemical air separation, oxygen pumping, and thermochemical heat pumping. A number of recent novel process developments are also presented, which offer valuable motivation and direction for the respective fields. The analysis shows that oxides which undergo a stoichiometric phase change during reduction, such as ZnO or Co3O4, have much larger specific energy storage than materials undergoing partial reduction such as ceria and perovskites. The partial reduction materials generally have faster kinetics and better activity at low temperature, and the selection of materials for the various applications is often a compromise between the importance of high specific energy storage vs.\ fast kinetics and low temperature operation.
Article
Full-text available
Splitting CO2 with a thermochemical redox cycle utilizes the entire solar spectrum and provides a favorable path to the synthesis of solar fuels at high rates and efficiencies. However, the temperature/pressure swing commonly applied between reduction and oxidation steps incurs irreversible energy losses and severe material stresses. Here, we experimentally demonstrate for the first time the single-step continuous splitting of CO2 into separate streams of CO and O2 under steady-state isothermal/isobaric conditions. This is accomplished using a solar-driven ceria membrane reactor conducting oxygen ions, electrons, and vacancies induced by the oxygen chemical potential gradient. Guided by the limitations imposed by thermodynamic equilibrium of CO2 thermolysis, we operated the solar reactor at 1,600°C, 3·10⁻⁶ bar pO2 and 3,500 suns radiation, yielding total selectivity of CO2 to CO + ½O2 with a conversion rate of 0.024 μmol·s⁻¹ per cm² membrane. The dynamics of the oxygen vacancy exchange, tracked by GC and XPS, further validated stable fuel production.
Article
Full-text available
Separation and concentration of O2 from gas mixtures is central to several sustainable energy technologies, such as solar-driven synthesis of liquid hydrocarbon fuels from CO2 , H2O, and concentrated sunlight. We introduce a rationale for designing metal oxide redox materials for oxygen separation through "thermochemical pumping" of O2 against a pO2 gradient with low-grade process heat. Electronic structure calculations show that the activity of O vacancies in metal oxides pinpoints the ideal oxygen exchange capacity of perovskites. Thermogravimetric analysis and high-temperature X-ray diffraction for SrCoO3-δ , BaCoO3-δ and BaMnO3-δ perovskites and Ag2O and Cu2O references confirm the predicted performance of SrCoO3-δ , which surpasses the performance of state-of-the-art Cu2O at these conditions with an oxygen exchange capacity of 44 mmol O2 mol SrCoO3-δ(-1) exchanged at 12.1 μmol O2 min(-1) g(-1) at 600-900 K. The redox trends are understood due to lattice expansion and electronic charge transfer. © 2015 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim.
Article
Most of the solar-driven thermochemical CO2 dissociation reactions have been focused on two-step processes. In this study, a one step CO2 thermolysis process was considered. It was found that direct thermolysis of CO2 obviously occurred at temperatures as low as 1200 °C within a corundum tube. The reaction rate could be enhanced by several times in the presence of metal oxides, which may be attributed to the catalysis of oxygen vacancies in the metal oxides.
Article
Solar powered two step thermochemical redox cycles are often investigated as a pathway to producing renewable fuels, via the splitting of water and carbon dioxide. These processes require a high temperature step for the reduction of the redox material. The reduction thermodynamics also critically depend on the oxygen partial pressure. In generic process assessments, purge gases and mechanical vacuum pumps have been considered for removing the released oxygen. Even though both alternatives are expected to have a significant impact on the overall process efficiency, little effort has been made so far in providing a detailed analysis of the implementation of the related components - especially in the case of vacuum pumps. In this study models are developed for assessing the energy requirements of mechanical vacuum pumps as well as for jet pumps. The models are validated and compared to data provided by manufacturers. In addition, a novel thermochemical pumping concept is introduced and a first thermodynamic assessment is presented. The energy demands of the different pump options are discussed in the context of the targeted application. By comparison with the energy stored in the produced fuel, limitations for the implementation of these systems can be identified. Furthermore, the costs for mechanical and jet pumps have been taken into account for the assessment. While mechanical pumps are energetically favored and could be reasonably applied for pressures down to 150 Pa, jet pumps are the more economic option and can make direct use of waste heat, which should be vastly available in the process at the required temperature level. The concept of thermochemical pumps is energetically very interesting and might shift the limit of realistic operational pressures to lower values, especially when combined with conventional systems.
Article
In this paper, ceria was proposed as a candidate material of membrane reactor for solar fuel production. A thermodynamic model of the membrane reactor system based on ceria with heat recovery was established and solar-to-fuel efficiency of both inert gas-assisted and pump-assisted CO2 splitting was calculated under a broad range of conditions. For system using inert gas, gas heat recovery is the determining factor for energy conversion efficiency. The energy efficiency is calculated to be >10% at 1800 K when the oxygen pressure at the inlet of reduction zone is lower than 10-6MPa. Increase of total pressure of the oxidation zone could improve the energy efficiency due to decrease of gas heat loss. Significant promotion in efficiency could be expected when a pump is applied to avoid using inert gas. Solar-to-fuel efficiency could be above 40% assuming good heat recovery. For the membrane reactor with a pump applied to maintain a vacuum atmosphere of the reduction zone, a simplified steady state model was put forward to predict the converting process and estimate the productivity. The diffusion rate of oxygen ions in the membrane is fast enough for conversion of considerable amount of CO2 in the reactor with a limited geometry.
Article
Production of chemical fuels by solar-driven thermochemical cycling has recently generated significant interest for its potential as a highly efficient method of storing solar energy. Of particular interest is the thermochemical process using non-stoichiometric oxides, such as ceria. In this process a reactive oxide is cyclically exposed to an inert gas, typically at 1500 °C to induce the partial reduction of the oxide, and then exposed to an oxidizing gas of either H_2O or CO_2 at a temperature between 800–1500 °C to oxidize the oxide and release H_2 or CO. Conventional wisdom has held that material kinetics limit the fuel production rates. Herein we demonstrate that, instead, at 1500 °C the rates of both reduction and oxidation of ceria, and hence also the global fuel production rate, are limited only by thermodynamic considerations for any reasonable set of operating conditions. Thus, in terms of materials design, significant room exists for sacrificing material kinetics in favor of thermodynamic characteristics.
Article
Relying more on the Sun Improved technologies for harnessing solar energy are not limited to creating more efficient solar cells. The associated hardware of delivering power from solar cells to homes and businesses, and storing this intermittent resource on the grid, offer R&D opportunities. Lewis reviews the status of these areas, as well as solar thermal and solar fuels approaches for harnessing solar energy. Science , this issue p. 10.1126/science.aad1920
Article
Thermodynamic analysis is performed on a conceptual oxygen permeation membrane reactor driven by concentrated solar energy (heat) for isothermal H2O splitting for the purpose of solar fuel derivation. By way of a plug-flow reactor model, kinetic and thermodynamic factors responsible for conversion rate, reactor dimension, and solar-to-fuel efficiency are analyzed for the case of pump-assisted and methane-assisted scenarios. The pump-assisted case achieves the same solar-to-fuel efficiency (2.9% at 1500 °C) as isothermal solar thermochemical cycling, while the methane-assisted case attains much higher efficiencies at much lower temperatures, whose net solar-to-fuel efficiency reaches 63% at around 900 °C. The theoretical framework developed in this study can be applied to the solar thermochemical splitting of other gases such as CO2 and can be further extended to the co-splitting of H2O and CO2 for syngas production driven by solar energy only (i.e., without the participation of hydrocarbon fuels). Copyright © 2015 John Wiley & Sons, Ltd.