You are on page 1of 11

Chemical Engineering Journal 346 (2018) 171–181

Contents lists available at ScienceDirect

Chemical Engineering Journal


journal homepage: www.elsevier.com/locate/cej

Assessment of kinetic model for ceria oxidation for chemical-looping CO2 T


dissociation

A.E. Farooquia,c, , A. Manfredi Picaa, P. Maroccoa, D. Ferreroa, A. Lanzinia, S. Fiorillib, J. Llorcac,
M. Santarellia
a
Energy Department (DENERG), Politecnico di Torino, Corso Duca degli Abruzzi 24, Torino 10129, Italy
b
Department of Applied Science and Technology (DISAT), Politecnico di Torino, Corso Duca degli Abruzzi 24, Torino 10129, Italy
c
Institute of Energy Technologies, Department of Chemical Engineering and Barcelona Research Center in Multiscale Science and Engineering, Universitat Politècnica de
Catalunya, EEBE, Eduard Maristany 10-14, Barcelona 08019, Spain

H I GH L IG H T S

• Experimental study of ceria oxidation by CO at 700–1000 °C and 20–40% CO .


2 2

• Maximum CO yield of 33.66 ml/g at 1000 °C and 40% CO . 2

• Statistical analysis for kinetic model selection.


• Models based on nucleation and grain growth reaction mechanism best fit the data.
• Sestak-Berggren model shows the best approximation of measured rates of reaction.

A R T I C LE I N FO A B S T R A C T

Keywords: Chemical looping technologies are identified as to have an excellent potential for CO2 capture and fuels
CO2 dissociation synthesis. Oxygen carriers are the fundamental component of a chemical looping process, and the choice of
Ceria stable and efficient carriers with fast redox kinetics is the key to the successful design of the process. Hence,
Kinetic model understanding the reaction kinetics is of paramount importance for the selection of an appropriate oxygen
Statistical analysis
carrier material. This work provides a method for kinetic model selection based on a statistical approach to
Experimental
identify the reaction mechanism. The study experimentally investigates the oxidation kinetics of CeO2−δ by CO2
and applies a statistical method for the selection of the best-fitting kinetic model for the reaction. The kinetic
study is performed in the temperature range of 700–1000 °C with a CO2 concentration between 20 and 40 vol%
in the feed. The measured peak rates of CO production on ceria were influenced both by temperature and
concentration of reactant. The total CO production was more influenced by the temperature than by CO2 con-
centration, with a maximum CO yield of 33.66 ml/g at 1000 °C and 40% CO2. The identification of the oxidation
kinetic model is performed by fitting different reactions models to the measured reaction rates and statistically
comparing them using the Residual sum of squares (RSS), Akaike information criterion (AICc) and the F-test for
the selection of the best-fitting one. Models corresponding to the nucleation and grain growth reaction me-
chanism provided a good fit of the data, with the Sestak-Berggren (SB) model showing the best approximation of
the measured rate of reaction with an evaluated activation energy of 79.1 ± 6.5 kJ/mol for the CO2 oxidation.

1. Introduction of the approaches to reduce CO2 emissions from fossil fuels is carbon
capture and sequestration (CCS). Carbon dioxide can be separated from
Global warming constitutes currently one of the most discussed the emissions originated from fossil fuels combustion by using different
environmental issues, and the scientific community converges toward technologies (e.g., absorption, adsorption, cryogenic distillation, etc.)
the awareness that the human impact on the environment is becoming and then sequestrated in geological formations or injected into nearly
more and more unsustainable. Most recognized responsible for such depleted oil/gas reservoirs for Enhanced Oil Recovery (EOR). Although
phenomenon are the greenhouse gases, mainly the carbon dioxide. One several CCS technologies exist, there are significant challenges still


Corresponding author at: Energy Department (DENERG), Politecnico di Torino, Corso Duca degli Abruzzi 24, Torino 10129, Italy.
E-mail address: azharuddin.xxx@polito.it (A.E. Farooqui).

https://doi.org/10.1016/j.cej.2018.04.041
Received 25 January 2018; Received in revised form 4 April 2018; Accepted 6 April 2018
Available online 07 April 2018
1385-8947/ © 2018 Elsevier B.V. All rights reserved.
A.E. Farooqui et al. Chemical Engineering Journal 346 (2018) 171–181

Nomenclature R gas constant that is equal to 8.314 J-mol/K


T temperature of the reaction (K)
Property Value A Arrhenius exponent
XCO,out mole fraction of CO at the exit of the reactor K reaction rate constant
ṅox ,out total molar outflow rate of the gas mixture g (α ) integral function of f (α )
P0 pressure at standard temperature and pressure (STP) αM conversion value at the maximum (dα/dt)
T0 temperature at standard temperature and pressure (STP) n avrami exponent
̇ 0 ,in
Vox total volumetric inflow rate at standard temperature and D grain size
pressure (STP) λ X-ray wavelength
mCeO2 mass of the ceria sample used. β line broadening at half the maximum intensity
δ1 non-stoichiometry reached after reduction step θ Braggs angle
δ2 non-stoichiometry reached after the oxidation of ceria step F F ratio
δ bulk-phase non-stoichiometry change RSS residual sum of squares
nO the accumulated intake of oxygen ions AIC Akaike information criterion
ω̇ CO CO rate (ml/min/g) x,y Sestak-Berggren (SB) model parameters
MCeO2 molecular weight of ceria β level of risk
nCeO2 moles of ceria γ maximum non-stoichiometry achievable for ceria
α conversion
Ea activation energy

associated with CCS, mostly due to safety and long-term stability and remains intact at the end of the cycle. Though the redox cyclic process
economic reasons [1]. The energy needs and environmental concerns, to produce syngas shows a promising potential, there are major chal-
thus, drive to look after alternative approaches, such as CO2 splitting lenges to achieve the high temperature required for the thermal re-
and/or utilization for the synthesis of fuels (e.g., methanol) [2–4]. duction, which needs a concentrated solar plant (CSP) [9]. Apart from
Solar-thermochemical dissociation of CO2 into CO attracted sig- high heat demand, the large temperature swing between the two-steps
nificant interest after the initial success of thermochemical H2O split- renders the process to be less efficient if not designed properly.
ting [5]. In fact, thermochemical cycles were initially focused on hy- Among many oxygen carrier materials for CO2 splitting, ceria has
drogen production from water splitting using oxygen carrier materials. been widely investigated in experimental studies. A number of experi-
In a solar-thermochemical cycle, the oxygen carrier participates in two ments [10–18] have demonstrated the feasibility of CO2 splitting with
separate redox reactions, in which it is first thermally reduced (using ceria, as listed in Table 1. However, only a few studies reported the
solar energy) and subsequently oxidized by H2O or CO2. The thermal reaction kinetics, mainly following equilibrium approach, defect model
reduction (TR) step of the cycle is endothermic and requires a higher theory, empirical solid state kinetics models [19–27]. Arifin [27] has
valence metal oxide, which releasing oxygen upon supply of external investigated the kinetics of splitting of water and CO2 over ceria and
heat forms a lower valence oxide of the metal. In the second step, the found that it is difficult to converge on a single kinetic model that
reduced metal oxide is oxidized back to higher valence state by taking adequately predicts the CO production behavior from thermally re-
oxygen from H2O or CO2 to form H2 or CO [6]. Here, a reduction duced ceria over the entire temperature range investigated. To achieve
temperature higher than the oxidation one is the thermodynamic con- a high quality fit to the data, three separate models had to be used
straint for this process to be attainable, as shown in Fig. 1. within the F family of models to give the best-fit to the CO transient
The net outcome of the cycle is the same as splitting H2O – or CO2 – signal with different kinetic parameters. Bulfin et al. [26] developed an
into H2 and O2 – or CO and O2 – in a single step reaction, but compared analytical kinetic model to fit experimental data and found that R3
to a single-step thermolysis reaction, it requires a sensibly lower tem- model gives the best fit results below 800 °C. Ackerman et al. [28] re-
perature (e.g., water thermolysis takes place at above 2300 °C) and it ported that D2 model provides the best-fitting for ceria oxidation at
also bypasses the problem of formation of explosive mixtures by pro- 1400 °C. The lack of agreement between the kinetic models based on
ducing separate streams of H2 or CO and O2 [7,8]. various experimental studies is a point of observation. The difference of
In this work, ceria-based chemical looping for CO2 splitting is in-
vestigated. Ceria has been chosen as it is considered one of the most
promising redox oxygen carriers because of its fast chemistry, high
ionic diffusivity and large oxygen storage capacity. Using oxygen-de-
ficient ceria, CO2 is dissociated into CO via:
Oxidation: CeO2 − δ + δCO2 (g) → CeO2 + δCO(g) (1)

Thermal reduction: CeO2 → CeO2 − δ + 0.5δO2 (g) (2)


where δ is the non-stoichiometric oxygen capacity, corresponding to the
surface oxygen vacancies determining the extent of CO2 dissociation
(Eq. (1)). Vacancies are formed back in the reduction step by the release
of oxygen (Eq. (2)). In solar-thermochemical CO2 splitting the oxygen
removal step can be achieved either by heating ceria to a high tem-
perature (∼1400 °C) using concentrated solar irradiation (Eq. (2)) or by
reducing the oxygen carrier using H2 (or even other fuels, e.g. CH4),
also called as reactive-chemical looping CO2 splitting (RCL-CDS).
Hydrogen reduction: CeO2 + δH2 (g) → CeO2 − δ + δ H2 O(g) (3)
The transfer of oxygen between the two redox steps exploits the Fig. 1. Schematic of solar-thermochemical CO2 splitting to produce syngas with
non-stoichiometric oxygen capacity of the ceria, and the oxygen carrier metal oxide looping between two-step redox cycles.

172
A.E. Farooqui et al. Chemical Engineering Journal 346 (2018) 171–181

Table 1 kinetic models based on different reaction mechanisms (i.e., reaction


Total CO production by CO2 splitting on CeO2 for thermally reduced and H2 order, geometrical contracting, diffusion, and nucleation models) are
reduced Ceria for various redox temperatures cited in the literature. compared using statistical criteria – Residual sum of squares (RSS),
Temp °C (Red/Ox) Total CO (ml/g) Feed CO2 (%) Reducer Refs. Akaike information criterion (AICc) and the F-test – and the best-fitting
model is selected and the corresponding ceria oxidation mechanism is
1500/800 6.28 50% Thermal [7] identified.
1400/1000 2.35 50% Thermal [10]
The experimental setup consists of a horizontal alumina tubular
1400/1000 2.24 50% Thermal [11]
1600/1000 4.91 60% Thermal [9] reactor, a control unit, a gas delivery system and a real-time gas ana-
1527/827 1.99 8.3% Thermal [6] lysis system with an online mass spectrometer. The system layout and
1200/850 0.83 25% Thermal [12] the details of the reactor are shown in Fig. 2. Three Bronkhorst EL-
1500/800 4.03 38.5% Thermal [13]
FLOW mass flow controllers (MFCs) are used for the gas flow control.
1500/1500 2.02 100% Thermal [14]
1400/1000 1.23 100% Thermal [15]
The reactor is made of an alumina tube positioned inside a tubular
1100/500 13.45 0.5–40% H2 [16] furnace (Lenton UK) that provides an isothermal environment up to
827/827 20.93 4% H2 [18] 1600 °C. As shown in Fig. 2, the reactor consists of an outer alumina
900/900 22.71 14.30% H2 [17] tube with 90 cm length, inner diameter (i.d) of 50 mm, an inner con-
700/700 4.17 14.30% H2 [17]
centric 12 mm o.d alumina tube and 10 mm i.d with 1 m length. The gas
700/700 7.57–9.19 20–40% 5%H2/Ar Present study
800/800 17.73–19.28 20–40% 5%H2/Ar Present study flow passes through the inner tube where the sample is placed in the
900/900 23.90–28.05 20–40% 5%H2/Ar Present study center supported with quartz wool. A quadrupole mass spectrometer
1000/1000 28.51–33.68 20–40% 5%H2/Ar Present study (QMS) (Hiden Analytical Inc.) is used to analyze the gas composition.
The QMS has a response time of less than 0.3 s and a wide bandwidth of
species detection capability.
the reaction mechanisms adopted could be a consequence of variations Commercial ceria powder from Alfa Aesar (99.95% purity) is used
in the experimental conditions or in the morphology of CeO2 samples. for the reaction study. The sample was crushed and sieved to 32 µm
Therefore, the present work aims to statistically analyze the solid-state before the tests. A 250 mg amount of ceria powder is embedded in
reaction kinetics models that describe the oxidation of non-stoichio- quartz wool and placed at the center of the inner alumina tube. The
metric ceria with CO2 by comparing their fitting goodness to a broad set total flow rate into the reactor during the oxidation step is maintained
of experimental measures. These reaction kinetic models are listed in constant at 120 N ml/min (GHSV 28,800 ml/g/h) and the CO2 mole
Table S1 (Supplementary data) with related detailed formula. fraction is varied between 20% and 40%, balance argon. A fixed mix-
In this work, CO2 dissociation over ceria is investigated by experi- ture of hydrogen and argon (5% H2 concentration) is used as a fuel for
ments and the measured reaction rates are used for kinetic models se- the reduction step, as the present study focuses on the oxidation step
lection based on a statistical approach to identify the involved reaction and the analysis of related kinetics (Eq. (4)). As the production of 1 mol
mechanism. of CO leads to the consumption of 1 mol of CO2, the total molar flow
rate throughout the control volume remains constant.
2. Experimental section Each experiment foresees a cycle of four steps. The first is the ceria
reduction step where the mixture of argon and hydrogen is sent for
2.1. Reaction activity testing 30 min to ensure complete reduction. Then, a purging stream of pure Ar
is fed for 10 min, to remove the H2 present in the fixed bed. The next
Isothermal redox cycles of CeO2 commercial powders are carried step is oxidation reaction where a mixture of Ar and different con-
out in a horizontal tubular reactor in the temperature range of centrations (20%, 30% and 40%) of CO2 is sent for 15 min. The final
700–1000 °C. H2 is used for the ceria reduction in order to explore the step is again the purging with pure Argon for 10 min. Isothermal redox
maximum non-stoichiometric capacity (δ) achieved at a set-point cycles were performed at 700, 800, 900 and 1000 °C. The measure of
temperature while using a different concentration of carbon dioxide in CO concentration in the outlet flow of the reactor allowed to extra-
the oxidation step. The temperature swing is thus replaced by iso- polate the reaction rate of the oxidation reaction:
thermal operation for developing the kinetics. The CO production
during the oxidation reaction is measured using an online mass spec-
trometer. Based on the reactivity data from the experiments, many

Ceria
Inner Alumina reactor Outer Alumina reactor

Vent
Check Valve
MS

Electric furnace
Condenser
Bypass
H2/Ar

CO2
Ar

Fig. 2. Experimental set-up for testing in micro-reactor configuration.

173
A.E. Farooqui et al. Chemical Engineering Journal 346 (2018) 171–181

̇ ,out
XCO,out nox ̇ 0 ,in
XCO,out P 0Vox 0.9λ
ω̇ CO = = D=
mCeO2 mCeO2 RT 0 β cosθ (7)
(4)

where XCO,out is the measured mole fraction of CO at the exit of the where D, λ, β, and θ are the grain size, the X-ray wavelength, the width
reactor, ṅox ,out is the total molar outflow rate of the gas mixture for the at half maximum intensity, and the Bragg angle respectively. Crystallite
̇ 0 ,in micrographs were obtained through a field emission scanning electron
oxidation, which is equal to the inlet molar flow, while P 0 , T 0 and Vox
microscope (FE-SEM, JSM7800F) at an accelerating voltage of 5 kV.
are the pressure, temperature and the total volumetric inflow rate at
standard temperature and pressure (STP). The reaction rates are nor-
malized by the total ceria sample mCeO2 – i.e. 250 mg – used in the 3. Reactivity results
measurement. The derivation assumes the quasi-steady-state and ne-
glects the accumulation or depletion effect in the control volume as the Firstly, the results of the tests have been investigated considering
residence time of the gases is negligible with respect to the character- the CO2 splitting performance in terms of CO production rate (N ml/
istic time of the redox conversion. min/g) and total CO yield (N ml/g). With the purpose to highlight the
The bulk-phase non-stoichiometry change of ceria has been eval- dependence of the performance of the oxygen carrier on temperature
uated from the extrapolated oxidation rate. The oxidation reaction can and reactant gas concentration, the tests were carried in different ex-
be rewritten as the following equation (Eq. (5)): perimental conditions, as described in the following sections.
1 1
CO2 (g) + CeO2 − δ1 → CO(g) + CeO2 − δ2 3.1. Effect of temperature
δ δ (5)

where δ1 is the non-stoichiometry reached after reduction step, δ2 is Fig. 3 shows the CO production rate as a function of temperature
non-stoichiometry reached after the oxidation of ceria step, and from 700 to 1000 °C. In each plot, the reaction rate exhibits a fast initial
δ = δ1−δ2 is the bulk-phase non-stoichiometry change, which is calcu- stage, followed by a decrease. During oxidation, the fast initial CO in-
lated according to the following equation: crease corresponds to the rapid oxygen vacancies ion incorporation.
nO (t) Both temperature and reactant concentration play a role in determining
δ(t) = the maximum rate. The peak rate varies nonlinearly with temperature
nCeO2 (6)
and for temperature lower than 700 °C, CO production is limited due to
t
where nO (t) = ∫ ω̇ CO dt is the accumulated intake of oxygen ions,
0 the low oxygen non-stoichiometric factor. The increase of temperature
nCeO2 = mCeO2 /MCeO2 is the moles of ceria used in the experiment, with to 800 °C showed significant enhancement of peak production rates,
MCeO2 its molar mass. nearly 1.5-fold higher. Further increase of temperature from 800 °C to
The non-stoichiometry is basically the amount of oxygen that the 1000 °C produces a less marked peak growth. All the peaks occur at
solid reactant can accept from the CO2, so, starting from a reduced around 15–30 s, and the peak duration increases with the temperature,
state, it decreases while the reaction occurs due to the ceria oxidation. due to the higher amount of available oxygen sites. After the peak, CO
At the end of the oxidation, oxygen vacancies are depleted and no more production rate decreases sharply at all the temperatures and ap-
oxygen is incorporated in the material. The maximum non-stoichio- proaches zero between 90 and 110 s. This behavior shows strong tem-
metry is affected by the temperature, with an increase of available perature dependence of the CO rate profile, which becomes higher and
vacancies when the ceria is reduced at higher temperature. wider at higher temperature, indicating a high activation barrier asso-
ciated with the CO2 splitting process [19].
2.2. Material characterization Fig. 4(a) emphasizes the observed behavior of the peak rates at
varying temperature for different CO2 molar fractions. The peak rates
To confirm the lattice structure of the samples before and after exhibit a profile with a fast increase around the temperature of 800 °C.
cycles, X-ray diffraction (XRD) was performed using a PANalytical Fig. 4(b) reports the total CO production during the oxidation step in
X’pert MPD Pro diffractometer with Ni-filtered Cu Kα irradiation (wa- the redox cycle. In all the cases, an almost linear rise of total CO pro-
velength 1.5406 Å). All samples were scanned in the 2θ range from 5° to duction is observed from 9 ml/g at 700 °C to 33 ml/g at 1000 °C. Fig. 4
80° with a step size of 0.2°/s. For a rough estimation on sintering effect, also shows that the effect of CO2 concentration on the total production
crystallite size of samples before and after test were calculated from the of CO is sensibly lower than that induced by temperature variation. The
Scherrer equation (Eq. (7)) based on the reflection with the highest strong temperature dependence is evident from the earlier studies
intensity. [14,16,25,29].

50 50 50
(a) (b) (c)
CO Peak rate (ml/min/g)

CO Peak rate (ml/min/g)


CO Peak rate (ml/min/g)

40 40 40

CO2: 20% CO2: 30% CO2: 40%


30 30 30
700oC 700oC 700oC
800oC 800oC 800oC
20 900oC 20 900oC 20 900oC
1000oC 1000oC 1000oC
10 10 10

0 0 0
0 40 80 120 0 40 80 120 0 40 80 120
Time (s) Time (s) Time (s)

Fig. 3. CO production rate (ml/min/g) during the oxidation step with (a) 20% mole fraction of CO2 in the feed, (b) 30% mole fraction of CO2 in the feed, (c) 40%
mole fraction of CO2 in the feed with Argon. Temperature is varied from 700 to 1000 °C.

174
A.E. Farooqui et al. Chemical Engineering Journal 346 (2018) 171–181

50 35 0.5
(a) (b)

Max Non-stoichiometry ( max)


CO Peak rate (ml/min/g)

Total CO production (ml/g)


40 28 0.4

30 21 0.3

20 14 0.2
CO2: 20%
CO2: 30%
CO2: 40%
10 CO2:20% 7 CO2: 20% 0.1
CO2:30% CO2: 30%
CO2:40% CO2: 40%
0 0 0.0
700 800 900 1000 700 800 900 1000
Temperature (oC) Temperature ( oC)
Fig. 4. (a) Peak CO rates as a function of temperature and CO2 mol fraction during oxidation and (b) total CO production and maximum non-stoichiometry δmax
attained as a function of temperature and CO2 mol fraction during oxidation.

Fig. 4(b) also shows the maximum non-stoichiometric factor (δmax) for temperature 700 °C is 0.06 for all the CO2 concentration range, and
of ceria reached during the oxidation step at different temperatures and only at higher temperature it varies slightly with CO2 fraction. It is
with different feed CO2 concentration. The concentration of oxygen evident from the results obtained that the maximum delta increased
vacancies in the ceria slightly increases with an increase of CO2 con- linearly with temperature, at variance with the effect of CO2 con-
centration in the feed, mainly at the higher temperatures. centration in the feed, which produced slight variation.
Fig. 5 compares the difference in profiles of the non-stoichiometry Fig. 6(a) shows the effect of concentration within 20–40% con-
(δ) as calculated in Eq. (6) during oxidation. The initial stage of oxi- centration (balance Argon) at 900 °C: peak rate position shifts from 30 s
dation ends within 20 s, but accounts for more than 70% of the overall δ to 15 s with an increase of CO2 concentration. For higher temperatures,
change, while the remaining oxidation leads to a lower change of non- this behavior is not observed and most of the peak rate positions were
stoichiometry. It is evident that the oxygen carrying capacity increases in the first 20–30 s range. Fig. 6(b) reports the CO peak rate for the
due to higher extent of non-stoichiometry achieved at higher tem- three CO2 concentration investigated for temperature of 900 °C and it
peratures. It can be noted that the non-stoichiometry increases from reflects that the CO peak rate increases linearly with CO2 concentration
0.07 to 0.21 in the 700–1000 °C temperature range for 20% CO2 mole in feed. The slope of the curve is higher for higher temperature showing
fraction, and a maximum of 0.25 is reached at 1000 °C for 30% CO2 a more pronounced influence for temperatures above 800 °C (see Sup-
mole fraction. Similar non-stoichiometry results for oxidation are re- plementary data). It can be seen that the total CO production with
ported elsewhere [16]. different CO2 concentration for a particular temperature is minimal and
a similar behavior has been observed for all temperatures. It is seen that
for CO2 concentrations of 30% and 40% the maximum non-stoichio-
3.2. Effect of CO2 concentration
metry (δmax) reaches the same value as seen in Fig. 4(b) and Fig. 5(b)
and (c). The relatively lower dependence on CO2 concentrations have
The effect of the oxidant concentration on kinetics and on conver-
been reported elsewhere [7,12,16,19,26,30].
sion time was also investigated. Fig. 3 shows that the reaction time
reduces with the increase of CO2 partial pressure in the feed. Higher
peak rates and reduced time to achieve peak are achieved with higher 3.3. Microstructural analysis
CO2 concentration. Similar profiles were reported by Zhou et al. [17].
Even if the conversion time reduces with the increase of CO2 mole XRD patterns of ceria before and after the reaction cycle, shown in
fraction in the feed, the growth of CO peak rate with CO2 concentration Fig. 7, revealed a Cubic fluorite structure in both cases. A minor con-
balances this effect and the total conversion remains same for that tamination of silica observed in the sample after cycle is ascribed to
particular point of interest, as shown in Fig. 4(b). For instance, the δmax residual quartz wool used to fix the bed in the reactor. Compared to

0.3 0.3 0.3


(a) (b) (c)
Non-stoichiometry
Non-stoichiometry
Non-stoichiometry

0.2 CO2: 20% 0.2 CO2: 30% 0.2 CO2: 40%

700oC 700oC 700oC


800oC 800oC 800oC
0.1 900oC 0.1 900oC 0.1 900oC
1000oC 1000oC 1000oC

0.0 0.0 0.0


0 40 80 120 0 40 80 120 0 40 80 120
time (s) time (s) time (s)
Fig. 5. Non stoichiometry during oxidation step with varying temperature from 700 to 1000 °C for (a) 20% mole fraction of CO2 (b) 30% mole fraction of CO2 and (c)
40% mole fraction of CO2.

175
A.E. Farooqui et al. Chemical Engineering Journal 346 (2018) 171–181

50 50 50

Total CO production (ml/g)


o
900oC

CO Peak rate (ml/min/g)


900 C (b)
(a)

CO Peak rate (ml/min/g)


40 CO2: 20% 40 40
CO2: 30%
CO2: 40%
30 30 30

20 20 20

10 10 10

0 0 0
0 40 80 120 20 30 40
time (s) CO2 % mole fraction

Fig. 6. (a) CO production rate versus time for varying concentration of CO2 in the feed for 900 °C (b) CO peak rate and total production as a function of CO2
concentration.

XRD patterns before cycling, the peaks appear more intense after cy- illustrated in Fig. S3 (Supplementary file). Generally, most of the
cling, which indicates a growth of crystalline grains during the high models listed in Table S1 contain one fitting parameter (rate constant,
temperature process. Average crystallite size of ceria before and after k). Some models with more than one parameter are more complex and
cycles were calculated from Scherrer equation based on the strongest allow a better fit for the kinetic data. These models are grouped in three
reflection peak (1 1 1) and resulted 11.3 and 12.1 μm, respectively. FE- groups based on the number of independent variables i.e., one para-
SEM images at high magnification of 25,000× show the coexistence of meter, two-parameters and three-parameters models. SB model is a
micron-sized particles decorated with much smaller ceria particles. The three fitting parameters model, while Avrami-Erofe’ev model with the
size of the larger particles is not significantly affected by the reaction, at exponent n, (AEn) and Random pore model (RPM) are two-parameter
variance with the smaller ceria particles which, due to sintering, appear model. All the other models listed in Table S1 are one-parameter
larger after reaction. This is clearly seen in related FE-SEM images and models.
is coherent with the larger average size obtained by Scherrer. The evaluation of kinetic model parameters includes isoconver-
sional and isothermal reaction analysis [38]. In the present study, the
working envelope of chemical looping process for narrow temperature
4. Kinetic study of ceria oxidation range isothermal method was chosen. As reported by Han et al. [39] the
intra-particle heat gradients can be assumed negligible and thus the
Generally, the reaction mechanism of solid state reactions is de- particle can be approximated to be isothermal.
scribed using reaction-order models (F), geometrical contraction In this paper, the reaction kinetics study is carried out by fitting
models (R), diffusion-limited models (D), nucleation models [31–34] different models to the experimental data shown in Section 3 in order to
(also called as Avrami-Erofe’ev models, AE), random pore model (RPM) identify the solid-state reaction kinetic mechanistic model. The metric
[35], and the Sestak-Berggren (SB) [36], and Prout-Tompkins models usually adopted for the comparison of reaction models with
(PT) [37]. The schematic description of the general kinetic models is

Fig. 7. XRD patterns of ceria before and after the reaction cycle. (b) FE-SEM images of fresh and (c) cycled ceria.

176
A.E. Farooqui et al. Chemical Engineering Journal 346 (2018) 171–181

experimental evidences is the reaction rate, measured in term of time In other terms, Eq. (12) is also expressed as:
profile of reactant conversion or product yield [17]. Three methods are
g (α ) = K ·t (13)
used to compare several solid-state reaction kinetic models against
isothermal experimental data. In particular, the methods are: (a) the fit The slope of the curve g (α ) vs t is the parameter K. The slope be-
quality of the transient conversion, (α vs. t); (b) the fit quality of the tween natural log of K vs 1/T (Eq. (14)) gives the activation energy as
transient time derivative of conversion (dα/dt vs. α) and (c) the Han- negative slope. The intercept will be ln(APm) where P is the partial
cock and Sharp Method or model free method [40]. The model free pressure of CO2. The reaction order ‘m’ was evaluated by plotting ln
method is used to verify the category of the kinetic model and is pre- (APm) vs. ln P (see Eq. (15)) and the slope would be the reaction order
sented in Supplementary data section. and the intercept would help in yielding the ‘A’ value.
Ea
4.1. Model fitting method lnK = − + lnA·P m
RT (14)

The kinetic study needs as input data the extent of reaction during ln[A·P m] = lnA + mlnP (15)
time, which can be derived from the cumulative of the CO produced as
The basic procedure here is to utilize the kinetic expressions of the
Eq. (8).
models reported in Table S1 to match the experimental data in the form
i−1
of dα/dt vs α and α vs t profiles by fitting the value of the K parameter,
cum (ω̇ CO,i) = ∑ (ω̇ CO,p) + ω̇ CO,i
and then select the models with the smallest residual sum of squares
p=1 (8)
(RSS) among candidate models with the same number of parameters
The extent of reaction (α) for each time instant is given by Eq. (9). [42]. Two parameters models (i.e., Avrami-Erofe’ev (AEn)) and three
parameters models (i.e., Sestak-Berggren (SB)) need also the evaluation
cum (ω̇ CO,i)
α (ti ) = of additional parameters.
cum (ω̇ CO,N ) (9)
For Avrami-Erofe’ev (AEn) model its exponent n need to be eval-
In other words, the extent of reaction at time ti is the ratio between uated. The validation of Avrami exponent (n) starts from the identifi-
i-th value of the cumulative and the final value of the cumulative. It cation of a particular value of conversion, αM, which is evaluated at the
implies that α varies from 0 to 1. These values are the experimental αs maximum dα/dt for each experiment. Then the parameter ‘n’ comes
that should be compared with the αs coming from the models. from the Eq. (16).
To obtain the kinetic model, a mathematical equation should be 1
developed. The kinetic expression for gas solid reaction can be ex- Avrami exponent n =
1 + ln(1−αM) (16)
pressed as Eq. (10) [41]:
For SB model, we need to identify the value of the two unknown

= k1·f (α )·[P ]m parameters x and y for evaluating f(α) and g(α) (see Table S1). A similar
dt (10)
procedure has been adopted, as to evaluate αM that comes at maximum
where α is the conversion, k1
= Aexp(−Ea/ RT ) and P is the partial dα/dt. For the calculation of the parameters x and y, Eq. (17) has been
pressure of CO2, m is the reaction order and f(α) is a function of α used.
depending on the reaction mechanism. The coefficients A and Ea are the αM
Arrhenius parameters; Ea being the activation energy and R is the gas p=
1−αM (17)
constant that is equal to 8.314 J/mol/K.
The first step of calculations was the fitting of the model with the where p = x / y , in which x and y are evaluated for each individual case.
raw data. For this purpose, Eq. (9) was transformed to Eq. (11):
4.2. Statistical methods for model discrimination

= K ·dt
f (α ) (11)
4.2.1. Statistical analysis of models
where K = k1·P m
is expressed in terms of partial pressure of CO2. The The comparison among the 19 different kinetics models listed in
integral of the reaction model is expressed by integrating Eq. (11). Table S1, beyond a graphical observation, is made using statistical
tools. The statistics takes in input two sets of data. The reference set is
α dα
g (α ) = ∫0 f (α ) (12)
composed by the experimental data of α or dα/dt. This one will be
compared with the sets composed by the data obtained with the

Table 2
RSS and AICc values for the 20% CO2 concentration (lowest in each kinetic model category).
Method Model 700 °C 800 °C 900 °C 1000 °C

RSS AICc RSS AICc RSS AICc RSS AICc

dα/dt-α F1.5 0.136506 −4796.81 0.114664 −5586.85 0.099529 −6189.02 0.096052 −6253.45
R2 0.020423 −5889.13 0.019425 −6735.57 0.016411 −7448.94 0.015734 −7525.22
D3 0.052743 −5343.6 0.045676 −6182.37 0.041149 −6806.41 0.03937 −6880.45
AE2 0.000588 −7928.65 0.000535 −9060.22 0.000394 −10055.2 0.000586 −9838.38
AEn 0.000264 −8387.92 0.000221 −9628.32 0.000155 −10707.2 0.000274 −10371
SB 0.000271 −8369.78 0.000209 −9665.28 8.94E−05 −11088.6 0.000143 −10826.9
PT 0.033293 −5608.14 0.026461 −6535.56 0.024524 −7168.17 0.026009 −7171.88

α-t F1.5 13.77583 −2143.59 14.89744 -2437.94 16.02001 −2637.29 16.04467 −2655.33
R2 2.669721 −3087.13 5.017568 −3142.03 4.96974 −3455.45 5.038927 −3469.53
D1 14.30321 −2121.98 15.7158 −2403.34 15.9027 −2642.43 16.0377 −2655.63
AE2 0.086214 −5061.04 0.18854 −5265.09 0.284452 −5454.98 0.346139 −5352.24
AEn 0.290962 −4359.64 0.01741 −6804.41 0.071244 −6420.72 0.354522 −5333.42
SB 0.257819 −4427.17 0.019789 −6719.56 0.05686 −6576.35 0.244802 −5591.75

177
A.E. Farooqui et al. Chemical Engineering Journal 346 (2018) 171–181

different models. Two statistical methods are employed in parallel to limited to AEn, AE2 and SB.
verify the agreement about the best fitting kinetic model: the Residual Fig. 8ii represents dα/dt vs. α and it shows the R2 model fits well for
Sum of Squares (RSS) and the Akaike Information Criterion (AICc) [42]. temperature above 800 °C only in the region (0.3 < α < 1.0). Models
Later, if the best-fitting models are characterized by a different com- AE2, AEn and SB were instead in close agreement with dα/dtexp values
plexity (i.e. different number of parameters) an F-test [43] allows to in all the conditions investigated. Thus, for (dα/dt vs. α) method only
select the best one, comparing the models two by two. Detailed ex- AEn, AE2 and SB models fit well with 0 < α < 1.0.
planation of Akaike Information Criterion is stated in Supplementary In conclusion, nucleation model (Avrami Erofe’ev) and SB model
data. were in close agreement with the αexp and dα/dtexp in all the conditions
The model is identified for the best possible accurately using RSS. investigated. These models belong to the same reaction mechanism
The two parameter (AEn) or three parameter models (SB) are expected except that AEn is based on two-parameters (k, n) and SB is an even
to exhibit better fits for the kinetic data in terms of smaller RSS. more complex model based on three-parameters (k, x, and y). In order
RSS and AICc values are tabulated for all three CO2 concentrations to identify the best-suited model describing the mechanism, the authors
for the temperature range (700–1000 °C) as given in Tables S2–S4 in the adopted the F-test method to distinguish between the three models
Supplementary data. For each category of models considered in the selected (AE2, AEn and SB) under the same category of the reaction
study, the model which showed the lowest RSS and AICc values for both mechanism which are nested within each other based on the number of
(α vs. t) and (dα/dt vs. α) was selected. F1.5 model has the lowest RSS parameters.
and AICc value in reaction order mechanisms with the reaction order
value of n = 0.91 (listed in Table S1). In geometrical contraction 4.2.2. F-test
models R2 model has lower RSS and AICc values with n = 1.11. In The F-test is the most common method adopted to determine the
Diffusion based reaction mechanism D1 with n = 0.62 had lowest AICc. statistically significant model between different model versions with
Two nucleation models (AE2 and AEn) were selected because showed varying number of parameters, which are termed as nested. It can
similar values of RSS and AICc. This is due to the fact that the values of happen that sometimes F-test and AICc differ in agreement in their
Avrami exponent (n) for AEn – calculated with Eq. (16) – were close to choice of the winner model [45]. The procedure is to select the model
somewhere around 1.9, which is almost the same value of the exponent with the smallest RSS among all the models with the same number of
of AE2. Similar values of n were also predicted from Hancock and Sharp fitting parameters, and then compare the relative value of the F-ratio,
method applied to AEn (described in Supplementary data). Thus, the which is evaluated according to the Eq. (18).
AEn and AE2 are of same category: the category of nucleation and grain
growth. RSSj2
Fj = 2
The results for the selected models are listed in Table 2 and are RSSmin (18)
plotted in Fig. 8 for 20% and 40% CO2 concentration. Similar behavior
for the 30% CO2 concentration is obtained, as can be seen in Supple- where RSSj is the Residual Sum of Squares of model j and RSSmin is the
mentary data (Figs. S5 and S6). minimum RSS between the two models compared. F-ratio follows the
Fig. 8i(a) shows that AEn and SB are the closest to αexp at 700 °C Fisher distribution with R and F degrees of freedom (which correspond
with RSS of 0.29 and 0.25, and AICc values of −4359.6 and −4427.1 to the number of parameters). When the models are nested with dif-
respectively. The graphical visualization shows that R2, AE2, AEn and ferent number of parameters, F-ratio is compared with Fcritical value
SB are fitting well with the experimental conversion (α) values for (α (which is the upper quantile of the Fisher distribution evaluated with
vs. t). For higher temperatures (Fig. 8i(b)–(d)), a good fitting of αexp is the different degree of freedoms). The F-ratio in comparing two nested
regression models (with different number of parameters) is the result of

1.0 1.0 1.0 1.0


i(a) i(b) i(c) i(d)
Conversion ( )

Conversion ( )

Conversion ( )
Conversion ( )

700oC 800oC 900oC 1000oC


F1.5 F1.5 F1.5 F1.5
0.5 0.5 0.5 0.5 R2
R2 R2 R2
D1 D1 D1 D1
AE2 AE2 AE2 AE2
AEn AEn AEn AEn
SB SB SB SB
Exp Exp Exp Exp
0.0 0.0 0.0 0.0
0 25 50 75 100 0 25 50 75 100 0 25 50 75 100 0 25 50 75 100 12
time (s) time (s) time (s) time (s)
0.04 0.04 0.04 0.04
700oC
F1.5
R2 800oC
F1.5
R2 900oC F1.5
R2
1000oC F1.5
R2
D3 D3 D3 D3
AE2 AE2 AE2 AE2
AEn AEn AEn AEn
d /dt

SB SB SB SB
d /dt

d /dt
d /dt

d /dtexp d /dtexp d /dtexp d /dtexp


0.02 0.02 0.02 0.02

ii(a) ii(b) ii(c) ii(d)


0.00 0.00 0.00 0.00
0.0 0.5 1.0 0.0 0.5 1.0 0.0 0.5 1.0 0.0 0.5 1.
Conversion ( ) Conversion ( ) Conversion( ) Conversion ( )
Fig. 8. Comparison of α vs. t between the experimental data and models with a concentration of 40% of CO2 at different temperatures; i(a) T = 700 °C, i(b)
T = 800 °C, i(c) T = 900 °C, i(d) T = 1000 °C; and dα/dt vs. α between the experimental data and models with a concentration of 20% of CO2 at different tem-
peratures; ii(a) T = 700 °C, ii(b) T = 800 °C, ii(c) T = 900 °C, ii(d) T = 1000 °C.

178
A.E. Farooqui et al. Chemical Engineering Journal 346 (2018) 171–181

a test where the null hypothesis is that all of the regression parameters represents the analytical method adopted to verify the activation en-
are equal to zero. In other words, the null hypothesis is that the model ergy, and it came out as 79.9 ± 4.1 kJ/mol which is in-line with the
has no predictive capability. Basically, the F-test compares the model kinetic methodology we adopted in our study.
with zero predictor variables, and decides whether the added coeffi-
cients (parameters) improved the model. 5. Conclusion
In order to incorporate the uncertainty, a confidence level defined
as 100(1 − β)% is selected. Fixing β, which is the level of significance This work presents a detailed kinetics study of CO2 splitting on
(it is usual to refer to 1 − β as level of confidence), the acceptance ceria. The time resolved kinetics were measured in a horizontal tubular
limits, called quantiles of order 1 − β can be extrapolated from the reactor at atmospheric pressure. The ceria sample was alternatively
table of the Fisher distribution cumulative probability function. exposed to 5%H2 in Argon mixture in the reduction step to remove the
Then, the Fcritical has been evaluated, which is the upper quantile lattice oxygen, and CO2 in the oxidations step to produce CO in the
calculated as: qf(1 − β, dfR − dfF, dfF) using F-statistic tables, where dfR redox cycle. Tests were performed under isothermal conditions
and dfF are the degrees of freedom of the model with a lower number of (700–1000 °C) for multiple redox cycles for three CO2 concentrations
parameters and a higher number of parameters, respectively. If F > qf between 20% and 40% in Argon. Experiments showed that with in-
(1 − β, dfR − dfF, dfF), then we reject the hypothesis and the model crease of temperature the total CO production increases. For instance
with the higher number of parameters is chosen as the best model [44]. the total CO production at 700 °C was 9.19 ml/g and peak production
After the identification of models which were with lowest RSS and was 13.25 ml/min/g, and for 1000 °C the total CO production was
AICc values, three models AE2, AEn and SB model were in agreement 28.15 ml/g and peak rate was 29.7 ml/min/g for CO2 concentration of
with experimental conversion values and F test with confidence interval 20%. For higher concentration of CO2 (40%) the total CO production
of 95% has been applied for both the methods adopted (α vs. t and dα/ increased to 33.66 ml/g and the peak rate to 46 ml/min/g for 1000 °C.
dt vs. α). The total CO production linearly increased with the increase of tem-
The F-test results for 20% CO2 concentration is listed in Table 3. It perature, and the effect of CO2 concentration on total production was
depicts that all F-values obtained from the best approximating models minimal. However, the effect of concentration of CO2 was seen with
at different temperatures are smaller than the critical value at 95% respect to peak rate signifying the time for conversion reduction.
confidence level (CL) (i.e., significant level β = 0.05). Therefore, the In order to identify the reaction mechanism and kinetic model,
model predictions and observed values are the same at 95% CL. F-test statistical approach was adopted to select among different reactions
results for 30% and 40% CO2 concentration (Tables S5 and S6) can be models by fitting them to experimental reaction rates. The activation
seen in Supplementary data. energies, the pre-exponential factors, and the reaction orders were de-
It is yielded that for temperature 700 °C, SB model was the winner termined. By determining the RSS and AICc values kinetic models were
for dα/dt vs. t method and AEn for α vs. t. Considering dα/dt vs. α selected, and based on the complexity and the number of parameters
method for all concentration of CO2 in the inlet, SB model is the winner the F-test considering upper quantile was used to select the best model.
model for temperatures 700 °C, 800 °C and 1000 °C except at 900 °C The results showed that the AEn and SB model are both suitable for
where AEn model is the winner. For α vs. t method, AEn model is the describing the oxidation reaction, but SB model was the best-fitting one
winner except at 800 °C with 20% CO2 and 1000 °C with 30% CO2, for most of the conditions of temperature and reactant concentration.
where SB is the winner model. The activation energy obtained considering the SB model was de-
Globally, the SB model is the winner in 32 of the F-tests, while AEn termined to be ∼79.1 ± 6.5 kJ/mol, in agreement with the literature.
in 27 and AE2 only in 13. Therefore, with the methodology adopted The present study gives a clear understanding in the model selection
both AE and SB model were in agreement with experimental data, but and mechanism of reaction for the entire range of conversion rate
AE2 and AEn passes the F-test in fewer conditions than SB, which re- (0 < α < 1.0), which would help in designing the chemical looping
vealed to be the winner model for the larger number of conditions in
either of the methods adopted. Though, it can be seen for few model Table 3
comparisons where the F values are very close to Fcritical, there may be F-test for the 20% CO2 concentration.
the chance that the winner model might change depending on the data.
Temperature Method Cases compared F-ratio Fcritical Winner model
It is observed only for 4 cases out of 24 cases reported in Table 3, such
as for SB/AE2 in conversion (α) vs. time. Choice of selection the model 700 dα/dt-t AEn/AE2 1.0113 1.1436 AE2
is subjective and is always taken along with RSS and AICc. SB/AE2 0.9627 0.8744 SB
SB/AEn 0.9519 0.8744 SB
α-t AEn/AE2 1.0170 1.1432 AE2
4.3. Kinetic parameter evaluation
SB/AE2 0.9224 0.8716 SB
SB/AEn 1.0198 1.1473 AEn
After the selection of SB as the best-fitting model, the kinetic
800 dα/dt-t AEn/AE2 1.0308 1.1341 AE2
parameters estimation is done. The ln(K) has been plotted versus in- SB/AE2 0.9595 0.8740 SB
verse of temperature (1/T) for each concentration of CO2 as described SB/AEn 0.9338 0.8740 SB
through equation 14. Fig. 9(a) represents the ln(K) vs (1/T) plot for the α-t AEn/AE2 1.0494 1.1334 AE2
three concentrations. The negative slope yields the activation energy SB/AE2 0.9215 0.8785 SB
SB/AEn 0.9674 0.8785 SB
for each concentration. The average activation energy evaluated from
the three concentrations is 79.1 ± 6.5 kJ/mol within a 95% confidence 900 dα/dt-t AEn/AE2 1.0060 1.1274 AE2
SB/AE2 1.0699 1.1274 AE2
level. The intercept of the value would be ln(APm), as described by Eq.
SB/AEn 1.0635 1.1274 AEn
(15), which is plotted in Fig. 9(b) against natural logarithm of partial α-t AEn/AE2 1.0098 1.1266 AE2
pressure of CO2. SB/AE2 0.9236 0.8828 SB
From Fig. 9(b) the reaction order obtained is 0.325 ± 0.06 and the SB/AEn 1.0107 1.1326 AEn
pre-exponential factor is 0.028 ± 0.01 s−1 Pam. The value of activation 1000 dα/dt-t AEn/AE2 0.9817 0.8893 AEn
energy is validated using an analytical model developed by Bulfin et al. SB/AE2 0.9644 0.8893 SB
[26], in which activation energy is obtained by plotting between ln(δ/ SB/AEn 0.9824 0.8893 SB
α-t Aen/AE2 0.9628 0.8894 AEn
(γ − δ)) vs 1/T. Here γ is the maximum non-stoichiometry that an
SB/AE2 0.9263 0.8831 SB
oxygen carrier can reach during oxidation and reduction step. For ceria SB/AEn 1.0636 1.1322 AEn
the value of γ reported as 0.35 [26]. Fig. S7 (Supplementary data)

179
A.E. Farooqui et al. Chemical Engineering Journal 346 (2018) 171–181

Fig. 9. (a) ln(K) vs (1/T) Arrhenius plot of the oxidation reaction for SB model, (b) ln(APm) vs ln(P) plot for oxidation reaction order determination.

CO2 splitting for packed bed or fluidized bed reactors for large scale [12] S.G. Rudisill, L.J. Venstrom, N.D. Petkovich, T. Quan, N. Hein, D.B. Boman,
system. J.H. Davidson, A. Stein, Enhanced oxidation kinetics in thermochemical cycling of
CeO2 through templated porosity, J. Phys. Chem. C 117 (2013) 1692–1700, http://
dx.doi.org/10.1021/jp309247c.
Acknowledgments [13] P. Furler, J. Scheffe, D. Marxer, M. Gorbar, A. Bonk, U. Vogt, A. Steinfeld,
Thermochemical CO2 splitting via redox cycling of ceria reticulated foam structures
with dual-scale porosities, Phys. Chem. Chem. Phys. 16 (2014) 10503–10511,
The research presented is performed within the framework of the http://dx.doi.org/10.1039/C4CP01172D.
Erasmus Mundus Joint Doctorate SELECT + program ‘Environomical [14] L.J. Venstrom, R.M. De Smith, Y. Hao, S.M. Haile, J.H. Davidson, Efficient splitting
Pathways for Sustainable Energy Systems’ and funded with support of CO2 in an isothermal redox cycle based on ceria, Energy Fuels 28 (2014)
2732–2742, http://dx.doi.org/10.1021/ef402492e.
from the Education, Audiovisual, and Culture Executive Agency
[15] B. Zhao, C. Huang, R. Ran, X. Wu, D. Weng, Two-step thermochemical looping using
(EACEA) of the European Commission. This publication reflects the modified ceria-based materials for splitting CO2, J. Mater. Sci. 51 (2016)
views only of the author(s), and the Commission cannot be held re- 2299–2306, http://dx.doi.org/10.1007/s10853-015-9534-7.
[16] S. Ackermann, L. Sauvin, R. Castiglioni, J.L.M. Rupp, J.R. Scheffe, A. Steinfeld,
sponsible for any use, which may be made of the information contained
Kinetics of CO2 reduction over nonstoichiometric ceria, J. Phys. Chem. C 119
therein. J. Llorca is a Serra Húnter Fellow and is grateful to ICREA (2015) 16452–16461, http://dx.doi.org/10.1021/acs.jpcc.5b03464.
Academia program and MINECO/FEDER project ENE2015-63969-R. [17] Z. Zhao, M. Uddi, N. Tsvetkov, B. Yildiz, A.F. Ghoniem, Enhanced intermediate-
temperature CO2 splitting using nonstoichiometric ceria and ceria–zirconia, Phys.
Chem. Chem. Phys. 19 (2017) 25774–25785, http://dx.doi.org/10.1039/
Appendix A. Supplementary data C7CP04789D.
[18] L.J. Venstrom, N. Petkovich, S. Rudisill, A. Stein, J.H. Davidson, The effects of
Supplementary data associated with this article can be found, in the morphology on the oxidation of ceria by water and carbon dioxide, J. Sol. Energy
Eng. 134 (2012) 11005, http://dx.doi.org/10.1115/1.4005119.
online version, at http://dx.doi.org/10.1016/j.cej.2018.04.041. [19] L. Zhu, Y. Lu, Reactivity and efficiency of ceria-based oxides for solar CO2 splitting
via isothermal and near-isothermal cycles, Energy Fuels (2017), http://dx.doi.org/
References 10.1021/acs.energyfuels.7b03284.
[20] N. Knoblauch, L. Dörrer, P. Fielitz, M. Schmücker, G. Borchardt, J.R. Scheffe,
A. Steinfeld, P.S. Manning, J.D. Sirman, J.A. Kilner, M. Kamiya, E. Shimada,
[1] D.Y.C. Leung, G. Caramanna, M.M. Maroto-Valer, An overview of current status of Y. Ikuma, M. Komatsu, H. Haneda, M. Katsuki, S. Wang, K. Yasumoto, M. Dokiya,
carbon dioxide capture and storage technologies, Renewable Sustainable Energy M. Stan, Y.T. Zhu, H. Jiang, E.N. Armstrong, K.L. Duncan, D.J. Oh, J.F. Weaver,
Rev. 39 (2014) 426–443, http://dx.doi.org/10.1016/j.rser.2014.07.093. E.D. Wachsman, B. Bulfin, A.J. Lowe, K.A. Keogh, B.E. Murphy, O. Lübben,
[2] Y.A. Daza, R.A. Kent, M.M. Yung, J.N. Kuhn, Carbon dioxide conversion by reverse S.A. Krasnikov, I.V. Shvets, S. Ackermann, J.R. Scheffe, A. Steinfeld, C.B. Gopal,
water-gas shift chemical looping on perovskite-type oxides, Ind. Eng. Chem. Res. 53 S.M. Haile, J. Rutman, M. Kilo, S. Weber, I. Riess, J. Maier, R.A. De Souza,
(2014) 5828–5837, http://dx.doi.org/10.1021/ie5002185. E.N. Armstrong, K.L. Duncan, E.D. Wachsman, H. Dünwald, C. Wagner, E. Fischer,
[3] J. Kim, C.A. Henao, T.A. Johnson, D.E. Dedrick, J.E. Miller, E.B. Stechel, J.L. Hertz, E.W. Hart, A.J. Mortlock, Q.-L. Meng, C.-I. Lee, T. Ishihara, H. Kaneko,
C.T. Maravelias, Methanol production from CO2 using solar-thermal energy: process Y. Tamaura, I. Ermanoski, N.P. Siegel, E.B. Stechel, R.J. Panlener, R.N. Blumenthal,
development and techno-economic analysis, Energy Environ. Sci. 4 (2011) 3122, J.E. Garnier, P. Fielitz, G. Borchardt, X.-D. Zhou, W. Huebner, I. Kosacki,
http://dx.doi.org/10.1039/c1ee01311d. H.U. Anderson, P. Fielitz, G. Borchardt, M. Schmücker, H. Schneider, P. Willich,
[4] S. Abanades, H.I. Villafan-Vidales, CO2 and H2O conversion to solar fuels via two- J.A. Lane, J.A. Kilner, M.A. Panhans, R.N. Blumenthal, C.C. Kan, H.H. Kan,
step solar thermochemical looping using iron oxide redox pair, Chem. Eng. J. 175 F.M.V.A.IV.E.N. Armstrong, E.D. Wachsman, Surface controlled reduction kinetics
(2011) 368–375, http://dx.doi.org/10.1016/j.cej.2011.09.124. of nominally undoped polycrystalline CeO2, Phys. Chem. Chem. Phys. 17 (2015)
[5] S. Abanades, G. Flamant, Thermochemical hydrogen production from a two-step 5849–5860, http://dx.doi.org/10.1039/C4CP05742B.
solar-driven water-splitting cycle based on cerium oxides, Sol. Energy 80 (2006) [21] Z. Zhao, M. Uddi, N. Tsvetkov, B. Yildiz, A.F. Ghoniem, Redox kinetics study of fuel
1611–1623, http://dx.doi.org/10.1016/j.solener.2005.12.005. reduced ceria for chemical-looping water splitting, J. Phys. Chem. C 120 (2016)
[6] P. Furler, J.R. Scheffe, A. Steinfeld, Syngas production by simultaneous splitting of 16271–16289, http://dx.doi.org/10.1021/acs.jpcc.6b01847.
H2O and CO2 via ceria redox reactions in a high-temperature solar reactor, Energy [22] E.V. Ramos-Fernandez, N.R. Shiju, G. Rothenberg, Understanding the solar-driven
Environ. Sci. 5 (2012) 6098–6103, http://dx.doi.org/10.1039/C1EE02620H. reduction of CO2 on doped ceria, RSC Adv. 4 (2014) 16456–16463, http://dx.doi.
[7] W.C. Chueh, C. Falter, M. Abbott, D. Scipio, P. Furler, S.M. Haile, A. Steinfeld, High- org/10.1039/C4RA01242A.
flux solar-driven thermochemical dissociation of CO2 and H2O using nonstoichio- [23] N. Knoblauch, H. Simon, L. Dörrer, D. Uxa, P. Fielitz, J. Wendelstorf, K.-H. Spitzer,
metric ceria, Science (80-) 330 (2010) 1797–1801, http://dx.doi.org/10.1126/ M. Schmücker, G. Borchardt, Ceria: recent results on dopant-induced surface phe-
science.1197834. nomena, Inorganics 5 (2017) 76, http://dx.doi.org/10.3390/inorganics5040076.
[8] P. Furler, J. Scheffe, M. Gorbar, L. Moes, U. Vogt, A. Steinfeld, Solar thermo- [24] M. Tou, R. Michalsky, A. Steinfeld, Solar-driven thermochemical splitting of CO2
chemical CO2 splitting utilizing a reticulated porous ceria redox system, Energy and in situ separation of CO and O2 across a ceria redox membrane reactor, Joule 1
Fuels 26 (2012) 7051–7059, http://dx.doi.org/10.1021/ef3013757. (2017) 146–154, http://dx.doi.org/10.1016/j.joule.2017.07.015.
[9] A. Stamatiou, P.G. Loutzenhiser, A. Steinfeld, Solar syngas production via H2 O/ [25] Z. Zhao, M. Uddi, N. Tsvetkov, B. Yildiz, A.F. Ghoniem, Redox Kinetics and
CO2-splitting thermochemical cycles with Zn/ZnO and FeO/Fe3O4 redox reactions, Nonstoichiometry of Ce0.5Zr0.5O2−δ for Water Splitting and Hydrogen Production,
Chem. Mater. 22 (2010) 851–859, http://dx.doi.org/10.1021/cm9016529. J. Phys. Chem. C 121 (21) (2017) 11055–11068.
[10] A. Le Gal, S. Abanades, G. Flamant, CO+ and H2O splitting for thermochemical [26] B. Bulfin, A.J. Lowe, K.A. Keogh, B.E. Murphy, O. Lübben, S.A. Krasnikov,
production of solar fuels using nonstoichiometric ceria and ceria/zirconia solid I.V. Shvets, Analytical model of CeO2 oxidation and reduction, J. Phys. Chem. C 117
solutions, Energy Fuels 25 (2011) 4836–4845, http://dx.doi.org/10.1021/ (2013) 24129–24137, http://dx.doi.org/10.1021/jp406578z.
ef200972r. [27] D. Arifin, Study of Redox Reactions to Split Water and Carbon Dioxide, University
[11] S. Abanades, A. Le Gal, CO2 splitting by thermo-chemical looping based on of Colorado, 2013, http://scholar.colorado.edu/chbe_gradetds/54.
ZrxCe1−xO2 oxygen carriers for synthetic fuel generation, Fuel 102 (2012) 180–186, [28] S. Ackermann, J.R. Scheffe, A. Steinfeld, Diffusion of oxygen in ceria at elevated
http://dx.doi.org/10.1016/j.fuel.2012.06.068. temperatures and its application to H2O/CO2 splitting thermochemical redox

180
A.E. Farooqui et al. Chemical Engineering Journal 346 (2018) 171–181

cycles, J. Phys. Chem. C 118 (2014) 5216–5225, http://dx.doi.org/10.1021/ org/10.1016/0040-6031(71)85051-7.


jp500755t. [37] F. Prout, E.G. Tompkins, The thermal decomposition of potassium permanganate,
[29] C. Muhich, A. Steinfeld, Principles of doping ceria for the solar thermochemical Trans. Faraday Soc. 40 (1944) 488–498.
redox splitting of H2O and CO2, J. Mater. Chem. A 5 (2017) 15578–15590, http:// [38] B. Janković, B. Adnađević, S. Mentus, The kinetic study of temperature-pro-
dx.doi.org/10.1039/C7TA04000H. grammed reduction of nickel oxide in hydrogen atmosphere, Chem. Eng. Sci. 63
[30] M. Hoes, C.L. Muhich, R. Jacot, G.R. Patzke, A. Steinfeld, Thermodynamics of (2008) 567–575, http://dx.doi.org/10.1016/j.ces.2007.09.043.
paired charge-compensating doped ceria with superior redox performance for solar [39] L. Han, Z. Zhou, G.M. Bollas, Heterogeneous modeling of chemical-looping com-
thermochemical splitting of H2O and CO2, J. Mater. Chem. A (2017), http://dx.doi. bustion. Part 1: reactor model, Chem. Eng. Sci. 104 (2013) 233–249, http://dx.doi.
org/10.1039/C7TA05824A. org/10.1016/j.ces.2013.09.021.
[31] M. Avrami, Kinetics of phase change. I general theory, J. Chem. Phys. 7 (1939) [40] J.D. Hancock, J.H. Sharp, Method of comparing solid-state kinetic data and its
1103–1112, http://dx.doi.org/10.1063/1.1750380. application to the decomposition of kaolinite, brucite, and BaCO3, J. Am. Ceram.
[32] M. Avrami, Kinetics of phase change. II transformation-time relations for random Soc. 55 (1972) 74–77, http://dx.doi.org/10.1111/j.1151-2916.1972.tb11213.x.
distribution of nuclei, J. Chem. Phys. 8 (1940) 212–224, http://dx.doi.org/10. [41] A.K. Galwey, M.E. Brown, Kinetic background to thermal analysis and calorimetry,
1063/1.1750631. Handb. Therm. Anal. Calorim. Vol. 1 Princ. Pract. 1998, pp. 147–216.
[33] M. Avrami, Granulation, phase change, and microstructure kinetics of phase [42] Kenneth P. Burnham, D.R. Anderson, Model Selection and Multi-model Inference: A
change. III, J. Chem. Phys. 9 (1941) 177–184, http://dx.doi.org/10.1063/1. Practical Information-theoretic Approach, second ed., Springer-Verlag, New York,
1750872. 2002.
[34] B.N. Erofe’ev, Generalized chemical kinetics equation and its application to reaction [43] M. Am, Introduction of the Theory of Statistics, third ed., McGraw-Hill, 1974.
with solid participants, Dokl. Akad. Nauk SSSR. 52 (1946) 515. [44] H.-V. Richard, G. Lomax, L. Debbie, Statistical Concepts, A Second Course, fourth
[35] S.K. Bhatia, D.D. Perlmutter, A random pore model for fluid-solid reactions: I. ed., Routledge, Taylot and Francis Group, New York, New York, 2012.
Isothermal, kinetic control, AIChE J. 26 (1980) 379–386, http://dx.doi.org/10. [45] T.M. Ludden, S.L. Beal, L.B. Sheiner, Comparison of the akaike information cri-
1002/aic.690260308. terion, the Schwarz criterion and the F test as guides to model selection, J.
[36] J. Šesták, G. Berggren, Study of the kinetics of the mechanism of solid-state reac- Pharmacokinet. Biopharm. 22 (1994) 431–445, http://dx.doi.org/10.1007/
tions at increasing temperatures, Thermochim. Acta 3 (1971) 1–12, http://dx.doi. BF02353864.

181

You might also like