You are on page 1of 11

Powder Technology 388 (2021) 474–484

Contents lists available at ScienceDirect

Powder Technology

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

Modulation of Fe-based oxygen carriers by low concentration doping


of Cu in chemical looping process: Reactivity and mechanism based
on experiments combined with DFT calculations
Nini Yuan, Hongcun Bai, Mei An, Jinpeng Zhang, Xiude Hu, Qingjie Guo 1,⁎
State Key Laboratory of High-efficiency Utilization of Coal and Green Chemical Engineering, College of Chemistry and Chemical Engineering, Ningxia University, Yinchuan 750021, China

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

Article history: Fe-based oxygen carriers (OCs) are currently the crucial material basis for chemical looping technology to realize
Received 21 January 2021 industrial applications. However, the relatively low reactivity of pure Fe-based OCs limits their extensive applica-
Received in revised form 8 April 2021 tions. One strategy to enhance reactivity in the chemical looping process is the introduction of another metallic
Accepted 9 April 2021
element by doping. This study prepared a series of Fe-based OCs (Cu2xFe2(1-x)O3) with low concentration doping
Available online 4 May 2021
of Cu. The reaction mechanism and reactivity modulation of Cu2xFe2(1-x)O3 OCs in the chemical looping process
Keywords:
were systematically investigated by means of experiments and density function theory (DFT) calculations. Acti-
Chemical looping vation energies for Cu2xFe2(1-x)O3 ranging between 72 and 37 kJ/mol were detected using H2 and the thermogra-
Oxygen carriers vimetric analyzer (TGA) test, indicating enhanced reactivity in the chemical looping process as compared with
Reaction mechanism that of pure Fe-based OCs of 84 kJ/mol. Thus, the low concentration doping of Cu can effectively improve the re-
Doping activity of Fe-based OCs. Furthermore, comprehensive DFT calculations upon the transition state indicated the re-
Density function theory calculations action energy barrier for Cu2xFe2(1-x)O3 with different doping concentration and configurations to be in the range
of 1.68–1.02 eV, lower than that of pure Fe2O3 of 2.30 eV. The Cu doping and the modulation of the reaction path-
ways are important reasons for the enhanced reactivity of Cu2xFe2(1-x)O3 OCs. Additionally, this study proposed a
lattice oxygen release mechanism of Cu2xFe2(1-x)O3 OCs during chemical looping combustion.
© 2021 Elsevier B.V. All rights reserved.

1. Introduction The design and preparation of high-efficiency OCs remains a popular


research topic and is highly relevant in the CLT field. The extensive re-
High-efficiency oxygen carriers (OCs) are currently a crucial mate- search of OC materials for large-scale application of CLT is mainly fo-
rial basis for chemical looping technology (CLT) to realize industrial ap- cused on Fe-based OCs [4,7,8]. This can be ascribed to the abundant
plications from laboratory-based studies. CLT [1,2] is a novel energy reserves, superior mechanical properties and environmental-
conversion technology, and its application can realize the capture of friendliness of Fe-based OCs. However, the reactivity of Fe-based OCs
CO2 [3] as feedstock to produce high value-added chemicals. Metal ox- is relatively low compared with those of Ni-based and Cu-based OCs,
ides are commonly employed as OCs [4,5] in CLT to transport lattice ox- which limits their wide applications. Several studies have investigated
ygen though redox reactions. Meanwhile, heat is transferred between the design and preparation of Fe-based OCs as well as their reactivities
an air reactor and a fuel reactor. CLT has the potential to realize the in the chemical looping process. For instance, Zhang et al. [9] studied
clean and efficient utilization of fuels and the reduction of pollutant the chemical looping gasification characteristics of rice straw with
emissions [6] without the need for an air-separating plant. The selection Fe2O3/Al2O3 OCs. The results of their study indicated that the addition
of high-efficiency OCs is a key requirement for CLT. These OCs are char- of inert Al2O3 improves the performance of Fe2O3 OC and the yield
acterized by a high oxygen-carrying capacity, high reactivity, and stable rate of syngas. Kuo et al. [10] prepared Na-modified Fe-based OCs
cycle performances. Therefore, the application of high-efficiency OCs using the conventional ceramic method and it was applied to chemical
can clearly reduce the quantity of OCs and equipment load required looping gasification process. The results indicated that the reduction
for the chemical looping system and can shorten the time required for rate of Fe-based OCs was enhanced by incorporation of Na. Additionally,
conversion of fuels in the reactor. Guo et al. [11] investigated that Fe2O3/ATP OC prepared by sol-gel
method was applied to the chemical looping combustion of coal. Their
⁎ Corresponding author.
results showed that the addition of inert ATP optimized the pore struc-
E-mail address: qingjie_guo@163.com (Q. Guo). ture of Fe-based OCs, increased their surface area, and improved the re-
1
Present address No.489, Helanshan west road, Xixia district, Yinchuan, Ningxia, China. activity and the conversion rate of coal. In an exploration of the

https://doi.org/10.1016/j.powtec.2021.04.022
0032-5910/© 2021 Elsevier B.V. All rights reserved.
N. Yuan, H. Bai, M. An et al. Powder Technology 388 (2021) 474–484

performance of an Fe2SiO2 OC with a core-shell structure, Bhavan et al. The results of the present study are expected to be helpful for the
[12] demonstrated a decrease in reactivity due to the formation of screening, design, and optimization of Fe-based OCs.
Fe2SiO4 and the loss of the core-shell structure during chemical looping
process. Although there have been significant efforts to investigate the 2. Experiment and computational method
composition, structure, and preparation of Fe-based OCs, single-
component Fe-based OCs continue to face the challenges of low reactiv- 2.1. Preparation of OCs
ity, rapid performance attenuation, and poor cycle performances during
the chemical looping process. The Cu2xFe2(1-x)O3 OCs (x = 1, 2, 5 and 10%) were synthetized via
Mono-metallic OCs such as Fe2O3, CuO, NiO, and Mn2O3 show differ- the co-precipitation method. Two kinds of nitrates (Cu(NO3)2·3H2O
ent advantages and disadvantages. Composite Fe-based OCs are consid- and Fe(NO3)3·9H2O) were dissolved in deionized water based on n
ered an alternative based on the complementary and synergistic effects (Cu2+)/n(Fe3+) = x/(1-x) to produce nitrate solution. The nitrate solu-
between binary metallic elements [13]. The purpose of the design and tion and the co-precipitation solution of 1 mol/L Na2CO3 were evenly
preparation of the Ni\\Fe [14], Cu\\Fe [15], Ca\\Fe [16,17], and Mn\\Fe mixed using a peristaltic pump with the pH of the mixture maintained
[18] bimetallic OCs is to produce the composite Fe-based OCs with ex- in range of 9–10. The precipitation mixture was then stirred for half
cellent comprehensive properties. For example, He et al. [19] prepared an hour until completely precipitated, the supernatant was discarded,
a Ni\\Fe bimetallic OC, which displayed a higher gasification efficiency and the material was washed repeatedly with deionized water until
for biomass due to the synergistic effects between Fe2O3 and NiO. How- neutral. Next, the material was vacuum filtered and transferred to an
ever, the Ni\\Fe alloy has been detected in the reduced samples, oven to dry at 120 °C for 12 h. Finally, the material was calcined in a
thereby changing the micro-morphology of the samples after 120 min muffle furnace at 900 °C for 3 h and then crushed and sieved using a
cyclic reactions and having a negative effect on lattice oxygen transport. standard sieve to obtain particles with a diameter between 75 and
Jiang et al. [20] investigated the dose-response effects of the addition of 150 μm. The blank sample of Fe2O3 OC was prepared using the same
CuO on the structure and properties of Cu\\Fe OCs. The addition of CuO method. All reagents used were of analytical purity and were provided
resulted in the OCs becoming porous and increased their reactivity, by the Chinese Medicines Group Chemical Reagent Co. Ltd.
whereas Cu\\Fe OCs were sintered with the addition of 20 wt% CuO.
Shen et al. [21] explored the metal synergistic effects of an CuFe2O4 2.2. Characterization
OC formed by mixing CuO and hematite. Their results indicated im-
provements to the reactivity of hematite and the physical stability of 2.2.1. XRD
CuO during the chemical looping process. Frick et al. [22] studied the The crystalline phases of OCs were characterized by X-ray diffraction
Cu\\Fe OCs by means of physical-chemical characterization, showing techniques (XRD, a Bruker D8 Advance A25) with a Cu target. The in-
that Cu\\Fe OCs of sufficient mechanical strength demonstrate high re- strument voltage, amperage, and scan rate were set to 40 kV, 40 mA,
activity with CO and CH4. Therefore, they were suitable for chemical and 2°/min from 5° to 85°, respectively.
looping combustion. Fan et al. [23] recently reported that 1% Cu-doped
modified Fe-based OCs was applied to the conversion of CH4 to produce 2.2.2. SEM-EDX
syngas during chemical looping gasification. The rate of CH4 conversion Scanning electron microscopy (SEM, ZEISS EVO18, Germany)
increased by 470% at 700 °C, and energy saving of ~35% when compared coupled with an energy dispersive X-ray system (EDX) mapping were
to that at 1000 °C. As a result, the selection of Cu as a dopant withing low used to analyze the micro-morphology and element distribution of
concentration modified Fe-based OCs shows obvious advantages. Fur- OCs. The instrument accelerated voltage and resolution were set to
ther investigation of Cu\\Fe OCs is important to advancing this technol- 10 kV and 1 μm, respectively.
ogy. Of particular importance is the study of the micro-reaction
mechanism and modulation of the reactivities of Cu\\Fe OCs, which 2.3. Reactivity test
only a few studies have reported on at the molecule/atom level.
The investigation of the micro-reaction mechanism of OCs during A thermogravimetric analyzer (TGA, Netzsch STA 449F3, Germany)
the chemical looping process [17,24] is helpful to understand the asso- was used to determine the conversion rate and lattice oxygen release
ciated structure-activity relationship and to provide theoretical guid- rate of OCs at various temperatures. H2 is an important gas component
ance for the design and preparation of OC materials. Density function that reacts with OCs during the chemical looping process. During the re-
theory (DFT) calculations are commonly used to provide valuable infor- activity tests, reducing gas containing 20 vol% H2 balanced Ar was used
mation on the micro-reaction mechanism, the active sites of the OC sur- for the evaluation of the reactivity of OCs at a gas flow rate of 200 mL/
face, and the adsorbability between OC and fuel molecules [18,24,25], min. The flow rate of the Ar protection gas was 20 mL/min and the
which is helpful to experimental studies [26,27]. For example, Liu heating rate was 10 °C/min.
et al. [28] investigated the reactivity and reaction mechanism of CO
over the CuFe2O4 OC surface and pointed out that the reaction of 2.4. Model and computational method
CuFe2O4 with CO was controlled by adsorption. Wang et al. [29] simi-
larly explored the solid-solid reaction of CuFe2O4 with C by the experi- DFT calculations were performed on the BAND module of the
ments and DFT calculations. However, there have been few reports on Amsterdam Density Functional (ADF) code [30–32]. In comparison
the modulation of reactivity of Fe-based OCs with low concentration with the plane-wave method adopted in the literature [23,33–35], the
doping of Cu. In particular, few studies have reported on the compre- basis set of the BAND module of the ADF code is the linear combination
hensive investigation of the micro-reaction mechanism and structure- of atomic orbitals (LCAO) and Slater-type orbitals (STOs). This repre-
activity relationship based on DFT calculations. sents an efficient and accurate approach for processing the system of
The present study investigated the micro-reaction mechanism and solids and periodic materials at an atom scale [36]. The DZP and TZP
modulation of reactivity of Cu2xFe2(1-x)O3 OCs (x = 1, 2, 5 and 10%) basis sets were tested, with the former providing accurate results for
with H2 during chemical looping combustion based on experiments the optimization of geometry, the calculation of reaction heat, and the
and DFT calculations. In addition, the detailed lattice oxygen release reaction mechanism investigated in the current study. The reaction
mechanism was explored and the composition, structure, element dis- mechanism of OCs and H2 was calculated by means of the exchange-
tribution, and micro-morphology of Cu2xFe2(1-x)O3 OCs were character- correlation function proposed by Perdew, Burke and Ernzerhof (PBE)
ized by X-ray powder diffraction (XRD) and scanning electron and a double-ζ plus polarization basis set DZP which approximates
microscopy/energy dispersive X-ray spectroscopy (SEM-EDS) mapping. the equivalent basis set 6-31G*, which was implemented in ADF code.

475
N. Yuan, H. Bai, M. An et al. Powder Technology 388 (2021) 474–484

A similar level of calculation was successfully used by Bai et al. [37,38] to chemical environment around Fe atoms, including changes to the type
process solid and periodic materials. The Nudged Elastic Band method of bond and an increase in the bond length. In addition, the changes to
(NEB) [25,39] was used to determine the transition state. The NEB the structural characteristics of nCu-Fe2O3 models (n = 2 and 3) were
method has been used in past studies to investigate the surface reaction similar to those of the Cu-Fe2O3 structure. The DFT calculations confirmed
mechanism of the periodic slab model [25]. The detailed calculations, in- obvious changes to the structure of Fe2O3 OC due to Cu doping.
cluding SCF convergence, force, displacement, and k-point sampling in
the first Brillouin zone were set as normal in the ADF program. 3.1.2. Adsorbability of H2 on the surface of OCs
As shown in Fig. 1, the α-Fe2O3 (001) surface with single iron-layer The present study explored the active sites of Cu2xFe2(1-x)O3 and the
(Fe-O3-Fe…termination) is the most stable and common among natu- stable adsorption configurations of H2 molecules on the surface of OCs
rally grown crystals [24,33,35,40], and was adopted to construct the based on the investigation of the adsorbability of H2 molecules on the
model of periodic slab. In the calculations, the super-cell model with 2 surface of OCs at different sites. In addition, the effects of the number
× 2 units contained 9 layer atoms. The setting of a 20 Å-thick vacuum and position of substituted Cu atoms on the adsorbability of H2 mole-
layer eliminated inter-slab-layer interactions. In addition, there was suf- cules were studied. As shown in Table 1 and Fig. 4, the present study ex-
ficient space for the relaxation of adsorption molecules. The bottom four plored a total of 30 adsorption models of H2 molecules on the surfaces of
layer atoms were fixed, whereas the remaining five layer atoms were nCu-Fe2O3, including three, six, nine, and twelve models for the surfaces
allowed to relax, constituting the slab model of Fe2O3. As shown in of Fe2O3, Cu-Fe2O3, 2Cu-Fe2O3, and 3Cu-Fe2O3, respectively.
Fig. 1(c), the single Fe atom of the Fe2O3 model surface was replaced The adsorbability of the surface of nCu-Fe2O3 was explored by focus-
by a Cu atom, following which the Cu-Fe2O3 model was generated. Fur- ing on the adsorption energy (ΔEads) and adsorption distance (D). D
thermore, the models with multiple substituted Cu atoms, primarily lo- represents the distance of a H2 molecule adsorbed on the OC surfaces.
cated the first, second, and third Fe atom layer of the surface and ΔEads is defined as below:
subsurface presented in the Fig. 1(b), were considered in the calcula- 
tions and labeled as nCu-Fe2O3 (n = 1, 2, 3 denotes the number of ΔEads ¼ EðOCþH2 Þ − EOC þ EH2 (1)
substituted Cu atoms). The calculated lattice parameters were a = b
= 5.038 Å and c = 13.750 Å, which is in good agreement with experi-
mental (a = b = 5.035 Å, c = 13.747 Å) and theoretical (a = b = In Eq. (1), E(OC+H2), EH2 and EOC represent the total energies of H2
5.035 Å, c = 13.74 Å) values [23]. The spin-polarization test was con- molecule adsorbed over the surface of OCs, the H2 molecule, and OCs,
ducted for the system of the Cu-Fe2O3 reaction with H2. The reaction en- respectively. Under this definition, the value of ΔEads is inversely related
ergy barriers were 1.69 eV and 1.68 eV for the top sites of the Cu atoms to the strength of interaction between H2 molecules and OCs.
with and without spin-polarization, respectively. Therefore, the basis As shown in Table 1, Eq. (1) was used to calculated the ΔEads of H2
set DZP without self-polarization was adopted for all calculations, molecule adsorbed over the surface of nCu-Fe2O3. The absolute values
which satisfied the basic calculation requirements for the current study. of ΔEads were all less than 0.2 eV. No new formations of chemical
bonds were observed. As shown in Fig. S1–S10 (from R-1 to R-30)
within the electronic supporting information (ESI), the process was
3. Results and discussion characterized by physical adsorption. As shown in Fig. 2, taking the
Fe2O3 OC as an example, the ΔEads of the top, bridge, and hollow sites
3.1. DFT calculations of the Fe atoms were − 0.13, −0.01, and − 0.07 eV, respectively,
whereas the adsorption distances corresponded to 2.150, 1.910, and
3.1.1. Structure analysis of OCs 1.780 Å, respectively. As shown in Fig. 2(c), these results indicated
The structures of OCs were investigated by the quantum chemistry that the strongest interactions between the H2 molecules and the
method. Fig. 1 shows the introduction of the Cu atom into the Fe2O3 struc- Fe2O3 OC occurred at the top site of the Fe atom. This site had the
ture taking the Cu-Fe2O3 model as an example. Partial Fe―O and partial smallest adsorption energy and the largest adsorption distance. This
Fe―Fe bonds with lengths of 1.754 Å and 3.150 Å respectively for the site also represented the configuration with the highest stability, similar
Fe2O3 model were replaced by Cu―O bonds and Cu―Fe bonds with to that of H2 adsorbed on the surface of iron oxide reported in literature
lengths of 1.882 Å and 3.190 Å, respectively. The increases in bond length [24,33,35,40]. As shown in Table 1, the adsorbability and adsorption
were as expected. In contrast to the Fe2O3 model, the structure of Cu- configurations of the H2 molecule on the surface of Cu-Fe2O3 were sim-
Fe2O3 not only contained an Cu atom, but also showed changes to the ilar to that of Fe2O3. The configurations of H2 molecules vertically

Fig. 1. Models of OCs: (a) primitive cell of α-Fe2O3; (b) α-Fe2O3 (001) surface; (c) Cu-Fe2O3 OC.

476
N. Yuan, H. Bai, M. An et al. Powder Technology 388 (2021) 474–484

Table 1
Adsorption distance (D), adsorption energy (ΔEads), reaction energy barrier (ΔEbarrier) and reaction heat (ΔEreaction) of H2 molecules on the surface of nCu-Fe2O3 oxygen carriers.

Models D④/Å ΔEads/eV ΔEbarrier/eV ΔEreaction/eV

Fe2O3-H2-top-Fe① 2.150 −0.13 2.30 −0.29


Fe2O3-H2-bridge-Fe 1.910 −0.01 2.47 −0.13
Fe2O3-H2-hollow-Fe 1.780 −0.07 2.64 −0.71

(Cu-Fe2O3)-H2-top-Cu② 2.140 −0.03 1.68 −0.34


(Cu-Fe2O3)-H2-bridge-Cu 1.870 −0.02 2.04 −0.36
(Cu-Fe2O3)-H2-hollow-Cu 1.530 −0.01 2.36 −0.43

(Cu-Fe2O3)-H2-top-Fe 2.180 −0.11 1.81 −0.36


(Cu-Fe2O3)-H2-bridge-Fe 1.920 −0.10 2.42 −0.46
(Cu-Fe2O3)-H2-hollow-Fe 1.700 −0.02 2.26 −0.60

(2Cu-Fe2O3)-1-adjoin-H2-top-Cu 2.086 −0.05 1.30 −0.32


(2Cu-Fe2O3)-1-adjoin-H2-bridge-Cu 1.565 −0.04 2.10 −0.28
(2Cu-Fe2O3)-1-adjoin-H2-hollow-Cu 1.568 −0.03 2.07 −0.32

(2Cu-Fe2O3)-1-diagonal-H2-top-Cu 2.118 −0.05 1.28 −0.55


(2Cu-Fe2O3)-1-diagonal-H2-bridge-Cu 1.438 −0.03 2.49 −0.55
(2Cu-Fe2O3)-1-diagonal-H2-hollow-Cu 1.531 −0.04 2.45 −0.56

(2Cu-Fe2O3)-1–2-H2-top-Cu 2.082 −0.07 1.29 −1.07


(2Cu-Fe2O3)-1–2-H2-bridge-Cu 1.760 −0.03 2.50 −1.09
(2Cu-Fe2O3)-1–2-H2-hollow-Cu 1.695 −0.06 2.32 −1.06

(3Cu-Fe2O3)-1-H2-top-Cu 2.057 −0.03 1.07 −0.48


(3Cu-Fe2O3)-1-H2-bridge-Cu 1.431 −0.01 1.28 −0.52
(3Cu-Fe2O3)-1-H2-hollow-Cu 1.557 −0.02 1.27 −0.48

(3Cu-Fe2O3)-1–2-H2-top-Cu 2.038 −0.04 1.02 −1.12


(3Cu-Fe2O3)-1–2-H2-bridge-Cu 1.657 −0.03 1.22 −1.12
(3Cu-Fe2O3)-1–2-H2-hollow-Cu 1.491 −0.03 1.42 −1.12

(3Cu-Fe2O3)-2–1-H2-top-Cu 2.058 −0.07 1.03 −0.58


(3Cu-Fe2O3)-2–1-H2-bridge-Cu 1.607 −0.04 1.71 −0.65
(3Cu-Fe2O3)-2–1-H2-hollow-Cu 1.592 −0.06 1.66 −0.60

(3Cu-Fe2O3)-1–2-3③-H2-top-Cu 2.064 −0.05 1.06 −1.03


(3Cu-Fe2O3)-1–2-3-H2-bridge-Cu 1.701 −0.03 1.22 −1.05
(3Cu-Fe2O3)-1–2-3-H2-hollow-Cu 1.709 −0.04 1.28 −1.03

Note:
① Fe2O3-H2-top (bridge/hollow)-Fe represents the adsorption and reaction of H2 molecules at the Fe-top (bridge/hollow) sites on the Fe2O3 surface.
② (Cu-Fe2O3)-H2-top (bridge/hollow)-Fe(Cu) represents the adsorption and reaction of H2 molecules at the Fe(Cu)-top (bridge/hollow) sites on the Cu-Fe2O3 surface.
③ 1,2,3 represents 1-Fe layer, 2-Fe layer and 3-Fe layer for nCu-Fe2O3 models, as shown in Fig. 1.
④ D represents the shortest distance of H2 molecule stable adsorbed on the OC surface.

Fig. 2. The stable adsorption configurations of the H2 molecule on Fe2O3 OC surface at different sites: (a) the top, bridge and hollow sites of Fe atom for Fe2O3 OC surface marked in yellow;
(b) the hollow site of Fe atom for over view; (c) the top site of Fe atom for side view; (d) the bridge site of Fe atom for side view.

477
N. Yuan, H. Bai, M. An et al. Powder Technology 388 (2021) 474–484

adsorbed on the top sites of metal atoms of the surface of Cu-Fe2O3 at the Cu atom sites preferentially occurred after adsorption following
showed higher stability and preferentially adsorbed forms. The metal the reaction at Fe atom sites. More importantly, the reaction pathways
atoms represented the active sites of the surface of the Cu-Fe2O3 OC. of Cu-Fe2O3 with H2 differed from those of Fe2O3 and H2 due to Cu doping.
As shown in Table 1 and Fig. S4–S10 (from R-4 to R-30) of the ESI, The reaction primarily originally occurred at the Fe atom sites for the one-
the adsorbability of H2 adsorbed on the Cu atoms of the surface of component Fe-based OC, whereas that for the Cu-Fe2O3 OC was at the Cu
nCu-Fe2O3 (n = 2 and 3) was investigated. The H2 adsorbed on the atom sites followed by the Fe atom sites. The reaction energy barriers of
top sites of Cu atoms on the surface of nCu-Fe2O3 showed a smaller the surface of Cu-Fe2O3 were lower than those of the one-component
ΔEads and more most distant D in comparison with those of the bridge Fe-based OCs due to Cu doping. In addition, obvious synergistic effects
and hollow sites, which were sites of higher stability and preferential of Cu-Fe were evident [13,20]. Analysis of the micro-reaction mechanism
adsorption. This result indicated that the number and position of demonstrated that the enhancement of the reactivity of Cu-Fe2O3 was
substituted Cu atoms had little effect on the adsorbability of H2 over mainly due to the existence of atomic-scale Cu doping defects.
Cu atom sites for the surface of nCu-Fe2O3. Fig. 3 shows the optimal reaction pathways. Within the reactant
The above results showed that the metal atoms of the surface of (R)-transition state (TS)-product (P) process, the H\\H bond of the H2
nCu-Fe2O3 were active sites and that H2 molecules were vertically molecule first stretches and disconnects. Taking Path 1 as an example,
adsorbed on the top sites of metal atoms through physical adsorption, H—H:0.751 → 1.204 → 1.590 Å. At the same time, the O atom of the sur-
representing the primarily stable adsorption configurations. face of Fe2O3, representing the nearest to adsorption sites of H2 molecule,
gradually move up. In addition, the Fe\\O bond continuously stretches
3.1.3. H2 oxidation on the surface of OCs and disconnects, path 1, Fe—O:1.832 → 2.081 → 2.580 Å. Subsequently,
The present study examined the reaction of H2 on the surface of it reacts with 2H atoms to form H2O: Path1, H—O:3.410 → 1.941 →
Cu2xFe2(1-x)O3 OCs during chemical looping combustion. The NEB 0.990 Å. Finally, the H2O molecule desorbs from the surface of the OC.
method was used to determine the transition states (TS) based on the The oxidation processes of H2 on the surfaces of Fe2O3 and Cu-Fe2O3
calculation conditions outlined in Section 2.4. The first step was the de- were similar. The H\\H bonds of the H2 molecule and metal‑oxygen
termination of the reaction start and end points. As shown in Fig. S1– bonds (Fe/Cu—O) of the surface of the OC disconnected to form H2O. Fi-
S10 (from R-1 to R-30) of the ESI, by regarding the start point of the re- nally, H2O desorbs from the surface of the OC. As shown in Fig. 3, the
action (R) to be the stable adsorption configurations of H2 molecules on main differences related to bond type and bond length between the
the surface of nCu-Fe2O3, the H\\H bond of H2 molecules was not dis- structure of reactant, transition state, and product. This result could be at-
connected and the system had the lowest energy. As shown in Fig. S1– tributed to the different structures of OCs as described in Section 3.1.1.
S10 (from P-1 to P-30) of the ESI, the stable configurations of the pro-
duced H2O molecules desorbed from the OC surfaces were taken as
the reaction (P) end point. The NEB method was used to generate five 3.1.4. Effects of the number and position of substituted Cu atoms on the re-
different images between the start and end points. The images with activity of OCs
the highest energy were considered as being in a transition state (TS), The present study investigated the micro-reaction mechanisms of
and their energies were used to calculate the energy barrier. Cu-Fe2O3, 2Cu-Fe2O3, and 3Cu-Fe2O3 with H2, respectively to explore
The reaction energy barrier (ΔEbarrier) was calculated to further un- the effects of the number and position of substituted Cu atoms on the re-
derstand the micro-reaction process of Cu2xFe2(1-x)O3 OCs with H2: activity of OCs. A reaction-mechanism analysis of Cu-Fe2O3 with H2
showed that the Cu atoms represented the surface activity sites and
ΔEbarrier ¼ ETS −ER (2) the optimal reaction sites in contrast to the Fe atoms. Therefore, as
shown in Table 1 and Fig. S1–S10 of the ESI, the focus was mainly on
In Eq. (2), as shown in Fig. S1–S10 (from TS-1 to TS-30) of the ESI, ETS the reaction of Cu atom sites on the surface of nCu-Fe2O3.
is the total energy of the transition state determined through NEB The 2Cu-Fe2O3 models with different Cu positions include three
method and ER is the overall energy of the reactant. Under this defini- kinds of configurations. As shown in Fig. S4–S6 of the ESI, the top,
tion, the value of ΔEbarrier is inversely related to the ease at which the re- bridge, and hollow sites of Cu atoms for each of the 2Cu-Fe2O3 models
action occurs and the reactivity of the OCs. were considered. Similar to the Cu-Fe2O3 model, the top sites of Cu
As shown in Fig. S1–S10 of the ESI, the present study explored the re- atoms showed the lowest ΔEbarrier and represented the optimal reaction
action of H2 on the different adsorption sites of the surface of nCu-Fe2O3. pathways. The ΔEbarrier was ~1.3 eV, indicating that the substituted po-
First, as shown in Table 1 and Fig. S1 of the ESI, the reaction of the Fe sition of Cu atoms had little effect on the reaction energy barriers for
atom sites on the surface of Fe2O3 was considered. The ΔEbarrier of the re- 2Cu-Fe2O3 models. As shown in Fig. 4, compared to ΔEbarrier(Fe2O3) =
actions of H2 on the top, bridge, and hollow sites of the Fe atom sites 2.30 eV, the ΔEbarrier of 2Cu-Fe2O3 surface decreased by ~1.0 eV. There-
were 2.30, 2.47, and 2.64 eV, respectively. The ΔEbarrier of the reaction fore, the reactivity of 2Cu-Fe2O3 was enhanced.
of H2 on the Fe atom top site was the lowest, indicating that the top As shown in Table 1 and Fig. S7–S10 of the ESI, similar to that pre-
site of the Fe atom was the most prone to chemical reactions and was as- sented for the 2Cu-Fe2O3 model, the present study investigated the reac-
sociated the optimal reaction pathway. This reaction pathway is Path 1 tion mechanism of 3Cu-Fe2O3 with H2. The lowest ΔEbarrier of ~1.0 eV
and is consistent with adsorbability of H2 on the surface of Fe2O3. occurred at the top sites of Cu atoms rather than at the bridge and hollow
Table 1 and Fig. S2–S3 of the ESI show the ΔEbarrier of the Fe and Cu sites. Similarly, the positions of substituted Cu atoms had little effect on
atoms sites of the surface of Cu-Fe2O3. The lowest energy barrier values the reaction energy barrier for the 3Cu-Fe2O3 models. As shown in Fig. 4,
at the Fe atom top site and Cu atom top site were 1.81 eV and 1.68 eV, the ΔEbarrier decreased by ~1.3 eV as compared to ΔEbarrier(Fe2O3) = 2.30
respectively. These also constituted the optimal reaction pathways for eV. As a result, the Cu atoms sites of the 3Cu-Fe2O3 surface exhibited a
the top site metal atoms of the surface of Cu-Fe2O3, namely Path 2 and superior reactivity compared to Fe2O3, Cu-Fe2O3, and 2Cu-Fe2O3.
Path 3. Fig. 3 shows the structures of reactant (R), transition state (TS), The above results indicated that the ΔEbarrier of H2 on the Cu atom
and product (P) in the three optimal reaction pathways. sites of the surface of nCu-Fe2O3 gradually decreased and the reactivity
As shown in Fig. 3, the ΔEbarrier of Fe2O3 reacted with H2 decrease from gradually enhanced as the number of substituted Cu atoms increased.
2.30 eV to 1.68 eV of Cu-top site and 1.81 eV of Fe-top site for Cu-Fe2O3 The oxidation of H2 molecules at the Cu atom sites of 3Cu-Fe2O3 oc-
surface. The decrease in the reaction energy barrier facilitated the reac- curred more readily. There were significant changes to the structure of
tion between Cu-Fe2O3 and H2. In addition, in contrast to Path 2 and the Fe2O3 OC when multiple Fe atoms (≤ 3 atoms) of the surface of
Path 3 of the surface of Cu-Fe2O3, the ΔEbarrier of the Cu atom sites were Fe2O3 were replaced by Cu atoms. As shown in Section 3.1.1, these
lower than those of the Fe atom sites. This result indicated the reaction changes included the introduction of active sites of Cu atoms in addition

478
N. Yuan, H. Bai, M. An et al. Powder Technology 388 (2021) 474–484

3.0
Path 1 Path 2 Path 3

2.5
2.30
H

2.0 R-1
Relative energy/eV

1.81
P-1
1.68
1.5
1.850
R-4
1.0 TS-1 P-4

0.5 R-7
TS-4
P-7

0.0 0.00
TS-7 -0.29
-0.36
-0.5 -0.34
R TS P
Reavtion coordinate

Fig. 3. The energy barrier diagram of OCs reaction with H2: Path 1: on Fe atom top sites of Fe2O3 surface; Path 2: on Fe atom top sites of Cu-Fe2O3 surface; Path 3: on Cu atom top sites of Cu-
Fe2O3 surface.

to those of Fe atoms, new chemical bonds, and an increase in bond


length. This was beneficial to H2 oxidation on the surface of nCu-Fe2O3. ΔEreaction ¼ Ep −ER (3)

In Eq. (3), Ep is the total energy of the produced H2O molecules


3.1.5. Reaction heat of OCs with H2 desorbed from the OC surfaces. A negative ΔEreaction value indicates an
The reaction heat of OCs with fuels in the reactor can be used to pro- exothermic reaction, whereas a positive ΔEreaction value indicates an en-
vide the thermal equilibrium required for the chemical looping system, dothermic reaction.
thereby reducing the required external heat supply and improving the As shown in Table 1, all ΔEreaction values of the reaction of nCu-Fe2O3
economy of the chemical looping process. Hence, there is value in ex- with H2 were negative, indicating that the reaction was exothermic,
ploring the change in heat occurring during reactions of OCs with H2. thereby supplying heat during the chemical looping process. Further-
The reaction heat (ΔEreaction) was calculated to investigate the more, the reaction heat increased from 0.29 eV for Fe2O3 to the maxi-
change in heat of the Cu2xFe2(1-x)O3 reaction with H2: mum 1.12 eV for nCu-Fe2O3 as the number of Cu atoms increased for

3
Fe2O3 Cu-Fe2O3 2Cu-Fe2O3 3Cu-Fe2O3
Fe sites Fe sites Cu sites Cu sites Cu sites
2.30
∆Ebarrier/eV

2
1.81

1.68

1.30

1.29
1.28

1.07

1.06
1.03
1.02

1
1 1 2 1 2 3 1 2 3 4
The number of the stable adsorption configrations of H2 molecule on nCu-Fe2O3 surface

top bridge hollow

Fig. 4. The energy barrier of different reaction sites for nCu-Fe2O3 OCs reacted with H2, respectively.

479
N. Yuan, H. Bai, M. An et al. Powder Technology 388 (2021) 474–484

nCu-Fe2O3. Therefore, Cu doping can modulate the heat produced


wi −wt
through the reaction between Fe2O3 OC and H2 during the chemical Ct ¼  100% (4)
wi −wx
looping process.

In Eq. (4), wi and wx represents the initial weight of the sample and
3.2. Experimental analysis the weight of the reduced OC, respectively. The rate of release of lattice
oxygen can be determined by the relationship between Ct and t. Finally,
3.2.1. Characterization of OCs the Arrhenius equation was used to calculate the apparent activation
As shown in Fig. 5, the XRD was used to characterize the composition energy, Ea.
and crystal structure of Cu2xFe2(1-x)O3 OCs. As shown in Fig. 5(a), all the Fig. 8 shows the Ct-t curves of the reaction between the Cu2xFe2(1-x)
characteristic diffraction peaks of Fe2O3 were present for both O3 OCs and H2. The value of Ct of the OC at unit time was positively re-
the fresh Fe2O3 and Cu2xFe2(1-x)O3 OC samples. This indicated that lated to the rate of release of lattice oxygen and the reactivity of the
the Cu2xFe2(1-x)O3 OCs were converted to Fe2O3 after calcination at OC. As shown in Fig. 8(a), the Ct of the Cu2xFe2(1-x)O3 OCs (x = 10%)
high temperature. No other impurity phase was evident. Furthermore, gradually increased with increasing temperature from 700 °C to 900
the doped Cu atoms completely entered the Fe2O3 lattice to replace °C. As shown in Fig. S-11 of the ESI, the trend of the conversion rate of
the equivalent number of Fe atoms [23]. As shown in Fig. 5(b), all lattice the Cu2xFe2(1-x)O3 OCs was the same as that of the Fe2O3. This result in-
oxygen removed from Fe2O3 OC was converted to metallic Fe, whereas dicated that the increase in temperature was beneficial to the rapid re-
the Cu2xFe2(1-x)O3 OCs were reduced to metallic Cu and Fe. lease of lattice oxygen of the Cu2xFe2(1-x)O3 OCs. More importantly, as
As shown in Fig. 6, the element distribution of the surface of the shown in Fig. 8(b), the Ct and rate of release of lattice oxygen of the
Cu2xFe2(1-x)O3 OCs was characterized through SEM-EDX mapping. The Cu2xFe2(1-x)O3 OCs exceeded those of Fe2O3 at the same temperature.
O, Cu, and Fe elements of the surfaces of Cu2xFe2(1-x)O3 (x = 10%) OCs The relationship of Ct and rate of release of lattice oxygen of the
were uniformly distributed with no local accumulation and agglomera- Cu2xFe2(1-x)O3 OCs at 700 °C was:
tion. This result indicated that a low concentration of Cu doping in the
Fe2O3 lattice could be realized through the co-precipitation method. In Cu0.20Fe1.80O3>Cu0.10Fe1.90O3>Cu0.04Fe1.96O3>Cu0.02Fe1.98O3>Fe2O3.
addition, the elements of the surfaces of the Cu2xFe2(1-x)O3 OCs were This result could be attributed to the Cu component increasing the
uniformly distributed. In particular, the original crystalline structure of reactivity of Cu-Fe OCs [42,43]. In addition, the Ct and rate of release of
Fe2O3 was maintained. Agglomeration and sintering did not readily lattice oxygen of the Cu2xFe2(1-x)O3 OCs gradually increased with in-
occur due to the amount and distributed forms of the Cu element creasing Cu content under a constant temperature. A low content of
[20,41]. Cu inhibited agglomeration and sintering during the chemical looping
As shown in Fig. 7, the morphologies of the Cu2xFe2(1-x)O3 (x = 10%) process [20]. Therefore, it was concluded that the use of Cu2xFe2(1-x)O3
OCs before and after the release of lattice oxygen were characterized by OCs is more suitable for the chemical looping process due to a higher
SEM. Fig. 7(a)(c) shows an increase in the particle diameter of the re- conversion rate and a rapid lattice oxygen release rate.
duced Fe2O3 OC. The shapes of particles were more obvious and gaps/ As shown in Fig. 9, the Ea of the reaction between the Cu2xFe2(1-x)O3
pores also appeared. The results shown in Fig. 7(b)(d) indicated in- OCs and H2 was calculated based on Ct-t curves and the Arrhenius equa-
creases in the particle diameter, pore size, and number of gaps of the re- tion. The Ea represents the energy required for the occurrence of the
duced Cu2xFe2(1-x)O3 OCs. In addition, the number of pores increased chemical reaction, with the value of Ea inversely related to the ease at
when the lattice oxygen was removed at 900 °C. Moreover, no agglom- which the chemical reaction occurs. Ea values of 83.9, 72.3, 66.0, 41.1,
eration and sintering was evident, which was due to the amount and and 37.0 kJ/mol corresponded with molar ratios of Cu doping of 0, 1,
distribution forms of the Cu element and the method of preparing the 2, 5, and 10%, respectively. This result indicated that the reactivities of
OCs [20]. Overall, doping using Cu under a low concentration changed Cu2xFe2(1-x)O3 OCs gradually increased with an increase doped Cu.
the structure of the Fe2O3 OC and optimized its micro-morphology. Low concentration Cu doping effectively improved the reactivity of
This was beneficial to the rapid release of lattice oxygen during the the Fe2O3 OC. The reactivity of the Cu2xFe2(1-x)O3 OCs gradually in-
chemical looping process. creased with increasing Cu content. More importantly, the concentra-
tion of doped Cu modulated the Ea and reactivity of the Fe2O3 OC,
3.2.2. Analysis of the reactivity of Cu2xFe2(1-x)O3 OCs consistent with the analysis of the mechanism of the reaction between
The conversion rate (Ct) was calculated by the TGA experiments to Cu2xFe2(1-x)O3 OCs and H2 shown in Section 3.1. As shown in
explore the reactivity of Cu2xFe2(1-x)O3 OCs. The sample weight wt at a Section 3.1.3, this result explained the microscopic nature of reactivity
given time t was converted to the Ct of OC as follows: enhancement of Cu2xFe2(1-x)O3 OCs.

(a) Fresh * * Fe2O3 (b) Reduced * Fe k Cu


* * *
* * *
Cu0.2Fe O
1.8 3
** Cu0.2Fe O
1.8 3 k * *
Intensity(a.u)
Intensity(a.u)

Cu0.10Fe O
1.90 3
Cu0.10Fe O
1.90 3
Cu0.04Fe O
1.96 3 Cu0.04Fe O
1.96 3
Cu0.02Fe O
1.98 3
Cu0.02Fe O
1.98 3
Fe2O3 Fe2O3

10 20 30 40 50 60 70 80 10 20 30 40 50 60 70 80
2θ (°) 2θ (°)

Fig. 5. XRD patters of Cu2xFe2(1-x)O3 OCs: (a) Fresh; (b) Reduced.

480
N. Yuan, H. Bai, M. An et al. Powder Technology 388 (2021) 474–484

Fig. 6. The SEM-Mapping images of the O, Cu and Fe elements distribution for Cu2xFe2(1-x)O3 OCs (x = 10%).

Fig. 7. SEM images of oxygen carrier: (a) Fresh Fe2O3; (b) Reduced Fe2O3 at 900 °C; (c) Fresh Cu2xFe2(1-x)O3 (x = 10%); (d) Reduced Cu2xFe2(1-x)O3 (x = 10%), at 900 °C.

481
N. Yuan, H. Bai, M. An et al. Powder Technology 388 (2021) 474–484

100 100
(a) (b) 700 ℃

Particle Conversion/%
Particle Conversion/%
80 80

60
Fe2O3
60
700 ℃ 1% Cu-Fe2O3
40 750 ℃ 40 2% Cu-Fe2O3
800 ℃
850 ℃
5% Cu-Fe2O3
20 900 ℃ 20 10% Cu-Fe2O3

0 0
0 10 20 30 40 50 0 10 20 30 40 50
Reaction Time /min Reaction Time /min

Fig. 8. The conversion rate curves of OCs at different temperature: (a) Cu0.20Fe1.80O; (b) at 700 °C, Cu2xFe2(1-x)O3 (x = 0, 1, 2, 5 and 10%).

2.5 3.3. Mechanism insights

The lattice oxygen release process of Cu2xFe2(1-x)O3 OCs can be di-


2.0
vided into different stages based on the TG and DTG curves shown in
Fig. 10. Fig. 10 (a) shows that the process of lattice oxygen release of
1.5 the Fe2O3 OC was divided into three stages. By combining with the re-
sults of XRD characterization shown in Fig. 5, the lattice oxygen release
ln (k)

process of Fe2O3 OC was:


1.0 Fe2O3 Ea=83.9 kJ/mol
Cu0.02Fe1.98O3 Ea=72.3 kJ/mol I:Fe2O3 → Fe3O4;II:Fe3O4 → FeO;III:FeO → Fe.

0.5 Cu0.04Fe1.96O3 Ea=66.0 kJ/mol


This is similar to what has been reported in the literature
Cu0.10Fe1.90O3 Ea=41.1 kJ/mol
[1,2,4,44,45]. As shown in Fig. 10(b)(c), the lattice oxygen release pro-
Cu0.20Fe1.80O3 Ea=37.0 kJ/mol cess for the Cu2xFe2(1-x)O3 OCs consisted of four stages. By combined
0.0
0.8 0.9 1.0 1.1 with result with the DFT calculations, two distinct reaction steps were
1/1000T (K-1)
identified. The first reaction step is the preferential reactions of Cu
Fig. 9. The activation energy of Cu2xFe2(1-x)O3 (x = 0, 1, 2, 5 and 10%) reacted with H2.
atom sites with lower reaction barriers, followed by the reactions of Fe
atom sites with relatively higher reaction barriers. Therefore, stage I

100 0.0 100 0.0


-0.5 -0.5
80 80
-1.0 -1.0
(a)
DTG/%/min

DTG/%/min
-1.5 (b) -1.5
TG/%

60 TG 60
TG/%

DTG TG
-2.0 DTG -2.0
40 -2.5 40 -2.5
-3.0 -3.0
20 20
䊠 䊢 -3.5 䊠䊡 䊢 䊣 -3.5

0 -4.0 0 -4.0
0 10 20 30 40 50 0 10 20 30 40 50
Reaction Time/min Reaction Time/min

100 0.0
-0.5
80
-1.0
DTG/%/min

(c) -1.5
TG/%

60
TG -2.0
DTG
40 -2.5
-3.0
20
-3.5
0
䊠䊡 䊢 䊣 -4.0
0 10 20 30 40 50
Reaction Time/min

Fig. 10. TG and DTG curves of OCs reacted with H2, at 700 °C: (a) Fe2O3; (b) Cu2xFe2(1-x)O3 (x = 2%); (c) Cu2xFe2(1-x)O3 (x = 10%).

482
N. Yuan, H. Bai, M. An et al. Powder Technology 388 (2021) 474–484

involved the reduction reaction of Cu atom sites and a higher release During the first step, the Cu2xFe2(1-x)O3 OCs were mainly converted
rate of lattice oxygen compared to those of the remaining stages of to metallic Cu and Fe2O3 at a rapid reaction rate. The second step consti-
Cu2xFe2(1-x)O3 OCs. The II, III, and IV stages of the Cu2xFe2(1-x)O3 OCs rep- tuted the stepwise reduction of Fe2O3 accompanied by the coexistence
resented the reduction reaction of the Fe2O3 OC with a low reaction rate. of metallic Cu reduced during the first step. The reactivities of the
As shown in Section 3.1, these results were consistent with the reaction Cu2xFe2(1-x)O3 OCs exceeded those of the Fe2O3 OCs. This result could
mechanism of Cu2xFe2(1-x)O3 OCs calculated by DFT. be attributed to the existence of atomic-scale Cu doped defects. The
TGA experiments and DFT calculations were used to investigate the bonding of the surface Cu atoms to O atoms (Cu\\O bonds) exhibited
reaction mechanism between Cu2xFe2(1-x)O3 OCs and H2 during chemi- higher activity. The energy barrier values of Cu atom sites were lower
cal looping combustion. As shown in Fig. 11, taking Cu-Fe2O3 as exam- than those of Fe sites. This result could be attributed to the high reactiv-
ple, the reaction of H2 on the surface of Cu-Fe2O3 included two distinct ity of Cu2xFe2(1-x)O3 OCs observed in the TGA experiments. In addition,
steps: (1) the oxidation of H2 molecules at Cu atom sites and the pro- although the reaction rate of the Cu2xFe2(1-x)O3 OCs clearly exceeded
duction of H2O required a minimum energy barrier of 1.68 eV; (2) the that of pure Fe-based OCs, the rate within the second step during the
oxidation of H2 molecules at Fe atom sites required a minimum energy chemical looping process remained lower, as observed in the experi-
barrier of 1.81 eV. However, the reaction of the single-component sur- ments and DFT calculations.
face of Fe2O3 was primarily the oxidation of H2 molecules at Fe atom
sites which required a minimum energy barrier of 2.30 eV. The analysis 4. Conclusions
of the mechanism of the microscopic reaction showed that the en-
hanced reactivity of Cu2xFe2(1-x)O3 OCs was mainly due to the existence The present study conducted a comprehensive investigation of the
of atomic-scale Cu doped defects. This significantly decreased the reac- reaction mechanism and modulation of reactivity between Cu2xFe2(1-x)
tion energy barrier of the reaction between Fe2O3 and H2 and modu- O3 OCs and H2 during chemical looping combustion based on TGA ex-
lated the reaction pathways. As shown in Section 3, the doped Cu periments and DFT calculations. The main conclusions are listed below.
atoms entered the Fe2O3 lattice and changed its structure, including
(1) The Cu2xFe2(1-x)O3 OCs with atomic scale Cu doped defects were
the slight shrinkage of the crystal cell from 5.036 to 5.035 Å, the intro-
prepared. The reactivity in chemical looping combustion was
duction of a new bond type and Cu surface active sites, and the increase
evaluated by the TGA experiments. The results showed that the
of bond length .
activation energy of Cu2xFe2(1-x)O3 OCs was in the range of
In contrast to the Fe2O3 OC, the structure of the Cu2xFe2(1-x)O3 OCs
72.3–37.0 kJ/mol, exhibiting a superior reactivity compared to
included not only Cu, Fe, and O atoms, but also a change to the chemical
that of Fe2O3 of 83.9 kJ/mol. DFT calculations confirmed that
environment around Fe atoms. Originally, the reaction primarily oc-
the reaction energy barrier was in the range of 1.68–1.02 eV
curred at Fe atom sites in the single-component Fe2O3 OC. The reaction
and lower than that of Fe2O3 of 2.30 eV and in good agreement
for the Cu2xFe2(1-x)O3 OCs occurred first at Cu atom sites, followed by
with higher reactivity of Cu2xFe2(1-x)O3 OCs observed in the
the Fe atom sites. In particular, the reaction energy barriers were
TGA experiments.
lower than those of the single-component Fe2O3 OC due to Cu doping.
(2) Systematic DFT calculations were conducted to reveal the micro-
Thus, the reactivity of the Cu2xFe2(1-x)O3 OCs were enhanced, consistent
scopic nature of reactivity enhancement for Cu2xFe2(1-x)O3 OCs.
with the results observed in the TGA experiments.
The reaction mechanisms of Cu2xFe2(1-x)O3 OCs with different
As discussed previously, the lattice oxygen release process of Cu2xFe2
doping concentrations and configurations were investigated.
(1-x)O3 OCs (x = 1, 2, 5 and 10%) included two distinct reaction steps, as
The results indicated that the energy barriers of Cu2xFe2(1-x)O3
follows:
OCs with H2 gradually decreased from 1.68 eV to 1.02 eV with
The first step: I:Cu2xFe2(1-x)O3 → Cu, Fe2O3;
an increase in the number of Cu atoms. The top sites of the Cu
The second step:II:Fe2O3 → Fe3O4; III:Fe3O4 → FeO; IV:FeO → Fe.
atoms were identified to be the optimal reaction sites. It was
found that Cu doping and the modulation of the reaction path-
ways constituted the primary reasons for the enhancement of
the reactivity of Cu2xFe2(1-x)O3 OCs based on the analysis of tran-
sition state structures and the reaction mechanism.
(3) The mechanisms of the release of lattice oxygen of Cu2xFe2(1-x)O3
OCs were studied. A model containing two distinct reaction steps
was proposed. During the first step, Cu2xFe2(1-x)O3 OCs were
mainly converted to metallic Cu and Fe2O3 at a rapid reaction
rate. The second step constituted the stepwise reduction of
Fe2O3 under a slow reaction rate. Although the reactivities of
Cu2xFe2(1-x)O3 OCs were superior to those of Fe2O3, the rates dur-
ing the second step of the chemical looping process remained
lower. The results of the present study can act as a reference for
the screening, design, and optimization of Fe-based OCs.

Declaration of Competing Interest

The authors declare that they have no known competing financial


interests or personal relationships that could have appeared to influ-
ence the work reported in this paper.

Acknowledgements

This work is mainly supported by National key research and devel-


Fig. 11. Schematic diagram of Cu-Fe2O3 oxygen carrier reaction with H2. opment program project (No. 2018YFB0605403-04), National Key

483
N. Yuan, H. Bai, M. An et al. Powder Technology 388 (2021) 474–484

R&D Projects of China (No. 2018YFB0605401), Key R&D Projects of [21] X. Niu, L. Shen, S. Jiang, H. Gu, J. Xiao, Combustion performance of sewage sludge in
chemical looping combustion with bimetallic Cu-Fe oxygen carrier, Chem. Eng. J.
Ningxia (No. 2018BCE01002), Natural Science Foundation of China 294 (2016) 185–192, https://doi.org/10.1016/j.cej.2017.01.091.
(No. 21868025), Financial supports from Innovation Program for Grad- [22] V. Frick, M. Rydén, H. Leion, Investigation of Cu-Fe and Mn-Ni oxides as oxygen car-
uate Students of Ningxia University (No. GIP2020054), Natural Science riers for chemical-looping combustion, Fuel Process. Technol. 150 (2016) 30–40,
https://doi.org/10.1016/j.fuproc.2016.04.032.
Foundation Project of Ningxia (No. 2020AAC03020) and National Aca-
[23] L. Qin, M. Guo, Y. Liu, Z. Cheng, J.A. Fan, L. Fan, Enhanced methane conversion in
demic Subjects Construction Project of Ningxia (Chemical Engineering chemical looping partial oxidation systems using a copper doping modification,
and Technology, grant No. NXYLXK2017A04). Appl. Catal. B Environ. 235 (2018) 143–149, https://doi.org/10.1016/j.apcatb.2018.
04.072.
[24] L. Huang, M. Tang, M. Fan, H. Cheng, Density functional theory study on the reaction
Appendix A. Supplementary data between hematite and methane during chemical looping process, Appl. Energy 159
(2015) 132–144, https://doi.org/10.1016/j.apenergy.2015.08.118.
Supplementary data to this article can be found online at https://doi. [25] F. Li, S. Luo, Z. Sun, X. Bao, L. Fan, Role of metal oxide support in redox reactions of
org/10.1016/j.powtec.2021.04.022. iron oxide for chemical looping applications: experiments and density functional
theory calculations, Energy Environ. Sci. 4 (2011) 3661–3667, https://doi.org/10.
1039/c1ee01325d.
References [26] V. Shah, Z. Cheng, D.S. Baser, J.A. Fan, L. Fan, Highly selective production of syngas
from chemical looping reforming of methane with CO2 utilization on MgO-
[1] L. Fan, L. Zeng, S. Luo, Chemical-looping technology platform, AICHE J. 61 (2015) supported calcium ferrite redox materials, Appl. Energy 282 (2021)
2–22, https://doi.org/10.1002/aic.14695. 116111–116123, https://doi.org/10.1016/j.apenergy.2020.116111.
[2] P. Wang, N. Means, D. Shekhawat, D. Berry, M. Massoudi, Chemical-looping combus- [27] W. Zhang, F. Zhang, L. Ma, J. Yang, J. Yang, H. Xiang, Reaction mechanism study of
tion and gasification of coals and oxygen carrier development: a brief review, Ener- new scheme using elemental sulfur for conversion of barite to barium sulfide, Pow-
gies 8 (2015) 10605–10635, https://doi.org/10.3390/en81010605. der Technol. 360 (2020) 1348–1354, https://doi.org/10.1016/j.powtec.2019.10.088.
[3] Th. Gauthier MYAF, CLC, a promising concept with challenging development issues, [28] F. Liu, J. Liu, Y. Yang, Z. Wang, C. Zheng, Reaction mechanism of spinel CuFe2O4 with
Powder Technol. (2017)https://doi.org/10.1016/j.powtec.2017.01.003. CO during chemical-looping combustion: an experimental and theoretical study, P.
[4] S. Luo, L. Zeng, L.S. Fan, Chemical looping technology: oxygen carrier characteristics, Combust. Inst. 37 (2019) 4399–4408, https://doi.org/10.1016/j.proci.2018.06.222.
Annu. Rev. Chem. Biomol. Eng. 6 (2015) 53–75, https://doi.org/10.1146/annurev- [29] T. Li, Q. Wu, W. Wang, Y. Xiao, C. Liu, F. Yang, Solid-solid reaction of CuFe2O4 with C
chembioeng-060713-040334. in chemical looping system: a comprehensive study, Fuel 267 (2020) 117163,
[5] L. Zeng, Z. Cheng, J.A. Fan, L. Fan, J. Gong, Metal oxide redox chemistry for chemical https://doi.org/10.1016/j.fuel.2020.117163.
looping processes, Nat. Rev. Chem. 2 (2018) 349–364, https://doi.org/10.1038/
[30] G. Velde, F M B E JGST, Chemistry with ADF, J. Comput. Chem. 22 (2001) 931–967,
s41570-018-0046-2.
https://doi.org/10.1002/jcc.1056.
[6] M. An, J. Ma, Q. Guo, Transformation and migration of mercury during chemical-
[31] V. Pershina, Relativistic effects on the properties of Lr: a periodic DFT study of the
looping gasification of coal, Ind. Eng. Chem. Res. 58 (2019) 20481–20490, https://
adsorption of Lr on surfaces of Ta in comparison with Lu and Tl, Inorg. Chem. 59
doi.org/10.1021/acs.iecr.9b04014.
(2020) 5490–5496, https://doi.org/10.1021/acs.inorgchem.0c00120.
[7] Z. Yu, Y. Yang, S. Yang, Q. Zhang, J. Zhao, Y. Fang, X. Hao, G. Guan, Iron-based oxygen
carriers in chemical looping conversions: a review, Carbon Resour. Convers. 2 [32] G. Te Velde, E.J. Baerends, Precise density-functional method for periodic structures,
(2019) 23–34, https://doi.org/10.1016/j.crcon.2018.11.004. Phys. Rev. B 44 (1991) 7888–7903, https://doi.org/10.1103/PhysRevB.44.7888.
[8] Z. Cheng, L. Qin, J.A. Fan, L. Fan, New insight into the development of oxygen carrier [33] C. Lin, W. Qin, C. Dong, H2S adsorption and decomposition on the gradually reduced
materials for chemical looping systems, Engineering 4 (2018) 343–351, https://doi. α-Fe2O3(001) surface: a DFT study, Appl. Surf. Sci. 387 (2016) 720–731, https://doi.
org/10.1016/j.eng.2018.05.002. org/10.1016/j.apsusc.2016.06.104.
[9] J. Hu, C. Li, Q. Zhang, Q. Guo, S. Zhao, W. Wang, D. Lee, Y. Yang, Using chemical [34] J.P. Perdew, K. Burke, Y. Wang, Generalized gradient approximation for the
looping gasification with Fe2O3/Al2O3 oxygen carrier to produce syngas (H2+CO) exchange-correlation hole of a many-electron system, Phys. Rev. B Condens. Matter
from rice straw, Int. J. Hydrogen. Energ. 44 (2019) 3382–3386, https://doi.org/10. 54 (1996) 16533–16539, https://doi.org/10.1103/physrevb.54.16533.
1016/j.ijhydene.2018.06.147. [35] C. Lin, W. Qin, C. Dong, Reduction effect of α-Fe2O3 on carbon deposition and CO ox-
[10] W. Huang, Y. Kuo, P. Su, Y. Tseng, H. Lee, Y. Ku, Redox performance of Na-modified idation during chemical-looping combustion, Chem. Eng. J. 301 (2016) 257–265,
Fe2O3/Al2O3 with syngas as reducing agent in chemical looping combustion process, https://doi.org/10.1016/j.cej.2016.04.136.
Chem. Eng. J. 334 (2018) 2079–2087, https://doi.org/10.1016/j.cej.2017.11.177. [36] R.A. Evarestov, Quantum Chemistry of Solids: The LCAO First Principles Treatment of
[11] Q. Guo, M. Yang, Y. Liu, Q. Yang, Y. Zhang, Multicycle investigation of a sol-gel de- Crystals, first ed. Springer Science & Business Media Press, Germany, 2007.
rived Fe2O3/ATP oxygen carrier for coal chemical looping combustion, AICHE J. 62 [37] H. Bai, W. Qiao, Y. Zhu, Y. Huang, Crystal orbital study on the combined carbon
(2016) 996–1006, https://doi.org/10.1002/aic.15100. nanowires constructed from linear carbon chains encapsulated in zigzag double-
[12] S. Bhavsar, M. Najera, G. Veser, Chemical looping dry reforming as novel, intensified walled carbon nanotubes, Curr. Appl. Phys. 15 (2015) 342–351, https://doi.org/10.
process for CO2 activation, Chem. Eng. Technol. 35 (2012) 1281–1290, https://doi. 1016/j.cap.2015.01.008.
org/10.1002/ceat.201100649. [38] H. Bai, Y. Zhu, W. Qiao, Y. Huang, Structures, stabilities and electronic properties of
[13] D. Zeng, Y. Qiu, S. Zhang, L. Ma, M. Li, D. Cui, J. Zeng, R. Xiao, Synergistic effects of bi- graphdiyne nanoribbons, RSC Adv. 1 (2011) 768–775, https://doi.org/10.1039/
nary oxygen carriers during chemical looping hydrogen production, Int. J. Hydrogen. c1ra00481f.
Energ. 44 (2019) 21290–21302, https://doi.org/10.1016/j.ijhydene.2019.06.118. [39] G. Henkelman, H. Jónsson, Improved tangent estimate in the nudged elastic band
[14] B. Wang, G. Xiao, X. Song, H. Zhao, C. Zheng, Chemical looping combustion of high- method for finding minimum energy paths and saddle points, J. Chem. Phys. 113
sulfur coal with NiFe2O4 combined oxygen carrier, J. Therm. Anal. Calorim. 118 (2000) 9978–9985, https://doi.org/10.1063/1.1323224.
(2014) 1593–1602, https://doi.org/10.1007/s10973-014-4074-y.
[40] X. Xiao, W. Qin, J. Wang, J. Li, C. Dong, Effect of surface FeS hybrid structure on the
[15] R. Siriwardane, H. Tian, J. Fisher, Production of pure hydrogen and synthesis gas with
activity of the perfect and reduced α-Fe2O3(001) for chemical looping combustion,
Cu-Fe oxygen carriers using combined processes of chemical looping combustion
Appl. Surf. Sci. 440 (2018) 29–34, https://doi.org/10.1016/j.apsusc.2017.12.252.
and methane decomposition/reforming, Int. J. Hydrogen. Energ. 40 (2015)
[41] L. Zhang, X. Wang, J.M. Millet, P.H. Matter, U.S. Ozkan, Investigation of highly active
1698–1708, https://doi.org/10.1016/j.ijhydene.2014.11.090.
Fe-Al-Cu catalysts for water-gas shift reaction, Appl. Catal. A Gen. 351 (2008) 1–8,
[16] G. Liu, Y. Liao, Y. Wu, X. Ma, Evaluation of Sr-substituted Ca2Fe2O5 as oxygen carrier
https://doi.org/10.1016/j.apcata.2008.08.019.
in microalgae chemical looping gasification, Fuel Process. Technol. 191 (2019)
93–103, https://doi.org/10.1016/j.fuproc.2019.03.019. [42] H. Leion, T. Mattisson, A. Lyngfelt, Using chemical-looping with oxygen uncoupling
[17] D.D. Miller, R. Siriwardane, CaFe2O4 oxygen carrier characterization during the par- (CLOU) for combustion of six different solid fuels, Energy Procedia 1 (2009)
tial oxidation of coal in the chemical looping gasification application, Appl. Energy 447–453, https://doi.org/10.1016/j.egypro.2009.01.060.
224 (2018) 708–716, https://doi.org/10.1016/j.apenergy.2018.05.035. [43] C. Rhodes, G.J. Hutchings, Studies of the role of the copper promoter in the iron
[18] F. Liu, J. Liu, Y. Yang, X. Wang, A mechanistic study of CO oxidation over spinel oxide/chromia high temperature water gas shift catalyst, Phys. Chem. Chem. Phys.
MnFe2O4 surface during chemical-looping combustion, Fuel 230 (2018) 410–417, 5 (2003) 2719, https://doi.org/10.1039/b303236c.
https://doi.org/10.1016/j.fuel.2018.05.079. [44] A. Pineau, N. KANARI, I. Gaballah, Kinetics of reduction of iron oxides by H2 part I:
[19] G. Wei, F. He, Z. Zhao, Z. Huang, A. Zheng, K. Zhao, H. Li, Performance of Fe-Ni bime- low temperature reduction of hematite, Thermochim. Acta 447 (2006) 89–100,
tallic oxygen carriers for chemical looping gasification of biomass in a 10 kWth in- https://doi.org/10.1016/j.tca.2007.01.014.
terconnected circulating fluidized bed reactor, Int. J. Hydrogen Energ. 40 (2015) [45] T. Song, L. Shen, J. Xiao, Z. Gao, H. Gu, S. Zhang, Characterization of hematite oxygen
16021–16032, https://doi.org/10.1016/j.ijhydene.2015.09.128. carrier in chemical-looping combustion at high reduction temperature, J. Fuel Chem.
[20] S. Jiang, L. Shen, J. Wu, J. Yan, T. Song, The investigations of hematite-CuO oxygen Technol. 39 (2011) 567–574, https://doi.org/10.1016/S1872-5813(11)60036-4.
carrier in chemical looping combustion, Chem. Eng. J. 317 (2017) 132–142.

484

You might also like