Next Article in Journal
A Review on the Synthesis of Polypeptoids
Next Article in Special Issue
Plasma-Enhanced Chemical Looping Oxidative Coupling of Methane through Synergy between Metal-Loaded Dielectric Particles and Non-Thermal Plasma
Previous Article in Journal
Set of Small Molecule Polyurethane (PU) Model Substrates: Ecotoxicity Evaluation and Identification of PU Degrading Biocatalysts
Previous Article in Special Issue
Ce-Doped LaMnO3 Redox Catalysts for Chemical Looping Oxidative Dehydrogenation of Ethane
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Recent Advances of Oxygen Carriers for Hydrogen Production via Chemical Looping Water-Splitting

1
School of Chemical Engineering, Northwest University, International Scientific and Technological Cooperation Base of MOST for Clean Utilization of Hydrocarbon Resources, Chemical Engineering Research Center for the Ministry of Education for Advance Use Technology of Shanbei Energy, Xi’an 710069, China
2
CAS Key Laboratory of Science and Technology on Applied Catalysis, Dalian Institute of Chemical Physics, Chinese Academy of Sciences, Dalian 116023, China
3
School of Chemical Engineering, University of Chinese Academy of Sciences, Beijing 100049, China
*
Authors to whom correspondence should be addressed.
Catalysts 2023, 13(2), 279; https://doi.org/10.3390/catal13020279
Submission received: 20 December 2022 / Revised: 12 January 2023 / Accepted: 19 January 2023 / Published: 26 January 2023

Abstract

:
Hydrogen is an important green energy source and chemical raw material for various industrial processes. At present, the major technique of hydrogen production is steam methane reforming (SMR), which suffers from high energy penalties and enormous CO2 emissions. As an alternative, chemical looping water-splitting (CLWS) technology represents an energy-efficient and environmentally friendly method for hydrogen production. The key to CLWS lies in the selection of suitable oxygen carriers (OCs) that hold outstanding sintering resistance, structural reversibility, and capability to release lattice oxygen and deoxygenate the steam for hydrogen generation. Described herein are the recent advances in designing OCs, including simple metal oxides (e.g., Fe, Zn, Ce, and Ti-based metal oxides) and composite metal oxides (e.g., perovskite, spinel, and garnets), for different CLWS processes with emphasis on the crucial parameters that determine their redox performance and future challenges.

1. Introduction

An urgent desire to reduce reliance on fossil fuels has been spurred in recent years due to the steeply increased concentration of greenhouse gases in the atmosphere [1]. Hydrogen, as an alternative, has been regarded as a promising fuel that possesses a high calorific value (121.00 MJ/kg) that is 2.8 times and four times greater than petroleum and coal, respectively, while giving clean water as the only combustion product [2,3,4]. More importantly, hydrogen is an important chemical raw material for various industrial processes, such as crude oil refining, synthesis of ammonia and methanol, etc. [5]. Currently, the production of hydrogen is mainly based on fossil fuels, such as steam methane reforming (SMR) and coal gasification (CG) [2,6,7]. However, both commercial routes are energy-intensive, rendering massive CO2 emissions reaching as high as 10–16 and 22–35 kgCO2e/kgH2 for the SMR and CG processes, respectively [8,9,10,11]. Although carbon capture and storage technologies can potentially reduce the CO2 emissions, this would inevitably induce a significant increase in energy penalty and manufacturing costs [12,13].
As a reverse reaction of hydrogen combustion, hydrogen production directly from water-splitting is considered a sustainable approach. Intuitively, direct thermal splitting of water (H2O(l) → H2(g) + 1/2O2(g), ΔHθ298 = 286 kJ/mol) is the simplest method to implement this goal. However, thermodynamic analysis shows that a positive Gibbs free energy is maintained for this reaction until a temperature higher than 4370 K (Figure 1), revealing an extremely tough process to break the O–H bond in H2O molecules [14]. Kogan et al. [15] proposed that water conversion only reached ca. 25% at a temperature of 2500 K and pressure of 0.05 bar, and rapid cooling treatment was needed to suppress the reverse reaction between generated H2 and O2, causing an increased energy burden. A study showed that hydrogen produced by pyrolyzing the mixture of water and argon was lower than 3% percentage of all off-gases at 2000~2500 °C [16]. Recent reports suggested that water-splitting driven by light and electricity enabled production of hydrogen at ambient temperature [17,18]. Compared with the photocatalytic process that is still in its infancy, electrochemical water-splitting could provide roughly 5% of the world’s hydrogen energy [19]. The purity of hydrogen generated in the electrolysis process is high, but the electrolysis process consumes a large amount of power, and the electricity cost accounts for around 80% of the overall water electrolysis process, rendering the scale-up of this technology strongly depending on the availability of low-cost and green electricity [20,21,22]. Therefore, research is ongoing to explore more efficient methods for hydrogen production with low carbon emissions and energy penalty.
Chemical looping technology that was initially designed as an advanced combustion method has shown great potential for energy storage and production of chemicals due to the virtue of process intensification, leading to improved process efficiency and notably reduced CO2 emission and investment expenses [23,24,25]. In recent years, pioneering reports have verified that applying the chemical looping concept to water-splitting, known as chemical looping water-splitting (CLWS), enables efficient hydrogen production at temperatures below 1300 °C. Compared with continuous reaction mode, CLWS reaction, utilizing metal oxides as oxygen carriers (OCs), is fulfilled by two or three spatiotemporal separated steps, including abstracting lattice oxygen from OCs via reduction treatment, regeneration by water with production of hydrogen, and in some cases, further oxidation by air to fully recover the OCs. Such a multistep reaction not only affords great convenience in separating hydrogen from products, but also requires less useful work and reaction free energy changes for water decomposition [26,27]. Furthermore, the introduction of OCs as redox catalysts significantly lowers the threshold to break the O–H bond of H2O and improves the reaction kinetics [28,29,30].
In general, the yield of hydrogen in a single redox cycle is positively correlated with the oxygen capacity of OCs, which in turn is closely related to the reduction method to remove the lattice oxygen. According to the distinct strategies in abstracting lattice oxygen from OCs, CLWS can be classified into processes including a two-step thermochemical water-splitting cycle (thermal reduction, Figure 2A) [31], a methane chemical looping process (CH4 reduction, Figure 2B), a chemical looping water gas shift reaction (CO reduction, Figure 2C) [32], a syngas (mixture of H2 and CO) chemical looping process (syngas reduction, Figure 2D) [33], and a photo-thermochemical cycle (photochemical reduction, Figure 2E) [34]. As for CLWS reaction, the economic efficiency of hydrogen production is substantially determined by redox properties of OCs, including mechanical strength, oxygen capacity, reactivity, selectivity, and structural stability, as well as low toxicity and low cost [35,36]. A series of reviews have been proposed in recent years that comprehensively summarized the potential applications of chemical looping technology [37], reaction mechanism of OCs [38], and the recent advantages in OCs for chemical looping conversion of alkanes [39,40,41]. As for hydrogen production via the chemical looping method, Voitic et al. [42] have reviewed the progress with emphasis on the iron-based OCs, while Luo et al. [43] detailed the system integration, economic analysis, and OCs for the chemical looping reforming process (cofeeding the fuel and water to reactor) and fuel-driven chemical looping water-splitting. Herein, the review gives a detailed survey on the current state of OC development for hydrogen production via different CLWS processes. The effect of the inherent properties of OCs on feed-gas conversion or product selectivity and future challenges in designing the OCs are summarized.

2. Processes for Chemical Looping Water-Splitting

2.1. Two-Step Thermochemical Water-Splitting

To address the problems of high reaction temperature and low conversion for direct water-splitting, in 1966 Funk et al. [26] proposed the strategy of two-step thermochemical water-splitting (TCWS) for hydrogen production. The key to this process lies in selecting suitable metal oxides as OCs, which are thermally reduced at high temperature (>1300 °C) to release oxygen and subsequently recovered by H2O at a temperature generally lower than 1100 °C with production of hydrogen. As a result, direct contact between hydrogen and oxygen is avoided, which bypasses the issue of hydrogen separation and enables higher water conversion at lower temperatures. In particular, when concentrated solar energy is utilized to heat the OCs (two-step solar thermochemical cycles), the solar energy is transformed to chemical energy with a theoretical solar-to-fuel efficiency reaching higher than 70% [14]. However, the high reaction temperature and thermal shocks induced by temperature swing pose a huge challenge in designing the OCs [44,45].
The Fe3O4/FeO redox pair was one of the earliest OCs used in a two-step TCWS cycle. Based on the thermodynamic analyses, Nakamura et al. [46] and Steinfeld et al. [47] reported that it was feasible to apply Fe3O4/FeO for hydrogen generation via thermochemical cycle at temperatures of ca. 2000 °C (Figure 3). However, these oxides were subjected to severe sintering under high operating temperatures. To address this issue, some active metals or inert supports were introduced to the OCs in an attempt to lower the thermal reduction temperature or improve their sintering resistance. Han et al. [48] found that removal of lattice oxygen from ferrites was realized at lower temperature of 1200 °C after doping of Ni, Mn, and Cu while water-splitting was implemented at 800 °C. Among the screened materials, NiFe2O4 displayed the highest rate for hydrogen production with an average H2 yield of 0.442 mL/g (the amount of produced H2 per gram of oxygen carrier in each cycle). Structural analysis revealed that no obvious changes occurred after these redox reactions, which demonstrated a high stability of this material. Kodama et al. [49,50] also suggested that Ni-modified ferrites in AxFe3−xO4 (A = Mn, Co, Ni) exhibited better reactivity than Co or Mn promoters, and a scaling relationship between Ni content (from 0 to 1) with corresponding performance was noticed. When Ni loading increased to x = 1, a maximum hydrogen productivity of ca. 10 mL/g is reached. Their follow-up studies showed that loading of ferrites on ZrO2 support significantly slowed down the sintering process, leading to stable hydrogen production. Ishihara et al. [51] reported that combination of NiFe2O4 with YSZ (yttrium-stabilized zirconia) support was capable of promoting the redox stability and oxygen mobility, resulting in a negligible decline in hydrogen production in 10 continuous cycles with reduction temperature of 1773 K and water-splitting temperature of 1473 K. Miller et al. [52] proposed that a mass ratio of 1/3 for Co0.67Fe2.33O4/YSZ could effectively inhibit the agglomeration of OCs and enable stable hydrogen production during eight cycles. In contrast, the bare Co0.67Fe2.33O4 oxides was inactivated in the second cycle.
In 1977, Bilgen’s team [16] applied a ZnO/Zn redox pair to the two-step TCWS cycle. Experimental results proved that thermal decomposition of ZnO into gaseous Zn (boiling point of 907 °C) and O2 was achieved at 2338 K using a solar furnace. Thermodynamic calculations showed that the theoretical solar-to-fuel efficiency (without heat recovery) of the water-splitting cycle could reach 51% when the redox cycle was conducted with thermal reduction temperature of 2338 K (ZnO decomposition) and water-splitting at 1500 K. However, practical application of ZnO in TCWS remains greatly challenging due to the facile recombination of Zn and O2 before introduction of water; thus, a quench treatment is necessary to separate Zn from O2 by rapid condensation of Zn gas into solid form. The numerical simulation results by Palumbo et al. [53] showed that a conversion of 80% was achieved for ZnO decomposition when a high solar flux density of 2 MW/m2 and quenching rate of 2.0 × 107 K/s were applied. Another challenge for application of ZnO/Zn redox pairs lies in the formation of a ZnO shell during the water-splitting reaction (Zn + H2O → ZnO + H2), which significantly slows down the reaction kinetics [54]. Therefore, recent works have been devoted to kinetic study on this process with the aim to improve the hydrogen production efficiency. Ernst et al. [55] investigated the reaction between submicron zinc and ~50 mol% H2O/Ar at 603–633 K. It was found that the initial reaction was a surface-controlled process with a fast rate and apparent activation energy of 42.8 ± 7.4 kJ/mol, while the subsequent reaction was a slow diffusion control process due to the restriction of surface ZnO layer on the migration of Zn2+. Lv et al. [56] investigated the effect of particle size on the water-splitting process and found that zinc powder with different sizes exhibited similar hydrolysis mechanisms, but that reducing the particle size could significantly enhance the reaction rate.
CeO2, as a benchmark OC for two-step TCWS, has attracted particular attention since 2006 due to its outstanding redox properties. Abanades et al. [57] found that Ce2O3 obtained by thermal reduction of CeO2 can be facilely oxidized by water with a conversion reaching 100%. However, a temperature of above 2000 °C is needed for complete reduction of CeO2 to Ce2O3, which inevitably resulted in serious sintering of OCs. To bypass this issue, in most cases for TCWS applications, the reduction temperature was lowered to below 2000 °C, leading to formation of non-stoichiometric CeO2−δ. Steinfeld et al. [58] reported that 2.5~2.9 mL of O2 can be released per gram of ceria in a single redox cycle (2.5~2.9 mL/g), corresponding to δ value of 0.039~0.044, when thermal reduction was conducted at ca. 1600 °C. This rendered a hydrogen evolution rate reaching a peak value of 7.6 × 102 mL/min during the following water-splitting step (<900 °C), which resulted in an overall solar-to-hydrogen efficiency of 0.7%. Costa Oliveira et al. [59] investigated the effect of CeO2 morphology on water-splitting reaction and discovered that H2 yield over three-dimensionally ordered macroporous (3DOM) CeO2 was approximately 32% greater than that of conventional CeO2 foam, suggesting that engineering the morphology was effective in adjusting the reaction kinetics and oxygen mobility. Although much progress has been achieved these years, the low oxygen capacity and high reduction temperature of pure ceria greatly limited the hydrogen productivity and process efficiency, which impeded the feasibility of ceria in practical STWS applications. To address this issue, significant research has been performed to tailor the redox properties of Ce-based OCs by tuning the oxygen vacancies and lattice strain via doping foreign cations [60,61,62,63]. Gokon et al. [64] compared the performance of 30 mol% M-doped CeO2 (M = Fe, Co, Ni, or Mn) in two-step TCWS cycles and found that all dopants except for Ni greatly increased the oxygen evolution during thermal reduction at 1500 °C. However, the yield of H2 is lower than that of bare CeO2, which may be ascribed to severe sintering of these OCs at high temperature. Tamaura et al. [65] investigated the performance of CeO2-MOx solid solution (M = Mn, Fe, Ni, Cu) with Ce/M molar ratio of 9 for two-step TCWS. As shown in Figure 4A, introducing of di- or tri-valent transition metal cations into CeO2 notably promoted H2 generation by enhancing releasing of lattice oxygen. The same group further screened a series of Ce0.9M0.1O2−δ ceramics (M = Mg, Ca, Sr, Sc, Y, Dy, Zr, and Hf), and found that the evolution of O2 (reduction of Ce4+ to Ce3+) in the thermal reduction step was improved by doping cations with higher valence state and smaller effective ionic radius (Figure 4B) [66,67]. Furthermore, the amount of H2 produced in the water-splitting step was dominated by the amount of O2 released in the reduction step. Instead, Abanades and coworkers [68] showed that addition of tri-valent dopants, including Y, La, Pr, and Gd, mainly contributed to improve the sintering resistance rather than the reducibility of ceria, but the ceria/zirconia mixed oxide was capable of greatly enhancing the reduction of bulk Ce4+. After detailed study on the correlation between zirconium content and redox properties of Ce1−δZrδO2 (0 ≤ δ ≤ 0.5), it was found that the ratio of Ce3+/Ce4+ in thermally reduced oxides gradually increased with Zr content while an optimal composition of Ce0.75Zr0.25O2 gave the highest hydrogen yield of 6.67 mL/g [69,70]. Their follow-up studies suggested that decoration of CeO2-ZrO2 with 1% Gd results in further promotion of hydrogen production to 7.58 mL/g [71]. Instead of exploring the nonstoichiometric chemistry that results in partial reduction of Ce4+ to Ce3+, Wang et al. [72] discovered that Sn-modified oxide of CeO2-xSnO2 (x = 0.05–0.20) was converted to thermodynamically stable Ce2Sn2O7 during the thermal reduction step (1400 °C), which enabled deep reduction of Ce4+ to Ce3+ rather than CeO2−δ. When the Sn content was optimized to x = 0.15, a hydrogen production as high as 7.18 mL/g (Figure 5), 3.8 times greater than that of unmodified CeO2, was obtained.
Perovskite oxides (ABO3) represent an important class of composite materials, wherein the A-site is generally occupied by rare earth, alkaline earth, or alkali metals that coordinate with 12 oxygen anions while the B-site is typically occupied by transition metals located in the octahedral interstices of the oxygen framework. These oxides hold virtue of outstanding redox properties, structure reversibility, and high-temperature stability, which render these oxides promising OCs for two-step TCWS reactions [73,74]. McDaniel et al. [75] discovered that Sr/Al doped LaMnO3 exhibited exceptional oxygen mobility, which released 8 times more O2 than CeO2 at 1350 °C, while the temperature for lattice oxygen desorption was nearly 300 °C lower than that of CeO2. Consequently, a hydrogen yield 9 times greater than that of CeO2 is obtained at 1000 °C. R. Barcellos and colleagues [76] demonstrated that doping of Ce at B-site of BaMnO3−δ allowed optimization of the oxygen vacancy formation energy. Among the investigated BaCexMn1−xO3−δ samples (0 ≤ x ≤ 1), BaCe0.25Mn0.75O3−δ displayed the best water-splitting performance and yielded nearly 3 times more hydrogen than ceria when reduced at 1350 °C. Similarly, Li et al. [31] also found that introduction of Co into LaGaO3 notably reduced the oxygen vacancy formation energy to ca. 3.0 eV, which was favorable for both lattice oxygen desorption and subsequent water-splitting. The best LaGa0.4Co0.6O3−δ sample exhibited a H2 yield of 0.478 mmol/g that was 15 times greater than CeO2 (0.032 mmol/g) when reduced at 1350 °C and re-oxidized at 800 °C. Carter et al. [77] screened a series of Ca0.5Ce0.5MO3 oxides (M = Sc, Ti, V, Cr, Mn, Fe, Co, and Ni) for a two-step solar thermochemical process using density functional theory (DFT) calculations. The calculation results suggested that the oxygen vacancy formation energy of Ca0.5Ce0.5MnO3 and Ca0.5Ce0.5FeO3 was slightly lower than that of CeO2 (4.0~4.3 eV) and located in the range of 3.65~3.96 and 3.77~4.06 eV, respectively, which fitted well to the target window (3.2~4.1 eV) for thermochemical cycles (Figure 6). Qian et al. [78] found that the CaTi0.5Mn0.5O3−δ was favorable for oxygen releasing and water-splitting due to its intermediate reduction enthalpy value of 200–280 KJ/mol-O (the standard enthalpy to remove one mole lattice oxygen of OC) and large entropy value of 120–180 J/mol-O·K (the standard entropy to remove one mole lattice oxygen of OC), which enables a hydrogen production of 10.0 mL/g in redox cycle between 1350 °C (thermal reduction) and 1150 °C (water-splitting) with a short cycle time of 1.5 h. All these results indicate that perovskite oxides are among the best candidates for two-step TCWS processes, and that A/B site engineering is efficient in modulating their redox properties. However, the step forward for practical application of this technology is still greatly hindered by high operating temperature and relatively low hydrogen yield due to the low oxygen redox capacity and long temperature swing time, which should be resolved ultimately by developing more effective OCs.

2.2. Methane Chemical Looping Process

The two-step TCWS cycle approach has substantially decreased the reaction temperature (≤1500 °C) compared to direct thermal splitting of water, but the high reaction temperature still puts forward tough requirements for the reaction equipment. Furthermore, the continuous temperature swing and thermal shock during the cyclic reaction not only reduces the reaction efficiency but also leads to rapid decrease of redox performance, which seriously limits the commercialization of this technology. To this end, recent reports verified that introduction of reducing gas could significantly accelerate the reduction kinetics of OCs, lower the reaction temperature to below 1000 °C, and increase the amount of lattice oxygen desorption, substantially enhancing the hydrogen production efficiency [79,80,81,82,83]. Methane, the main component of natural gas, has been widely studied due to its abundant reserves, high hydrogen-to-carbon ratio, and strong reducibility [84,85]. The two-step TCWS cycle with methane as reducing gas is commonly referred to as methane chemical looping process (Figure 2B). The facile reduction of OCs in methane atmosphere enables a closed reaction loop at much milder conditions, which potentially reduces the energy penalty and improves the economics of this process. Compared with conventional standalone systems that generate hydrogen and electricity separately based on methane, it is assessed that chemical looping hydrogen generation with methane as reduction gas was able to save more than 16% of energy input while reducing beyond 98% of CO2 emissions, rendering a low cost of $32.87/MWh for H2 production [86].
The methane conversion over OCs undergoes different processes, including partial oxidation or total combustion, when modulating the redox properties of these oxides. As for OCs with relatively low reducibility, syngas production is favored by selective methane oxidation (CH4 + OL → CO + 2 H2) (Figure 2B1). The reduced OCs can be subsequently recovered by water oxidation with generation of hydrogen (H2O → OL + H2). The overall reaction is generally referred as chemical looping steam methane reforming (CL-SMR) [87]. Compared with traditional steam methane reforming reaction (CH4 + H2O → CO + 3 H2), the CL-SMR process can obtain syngas with H2/CO ratio of 2, which is suitable for Fischer–Tropsch synthesis and methanol. Furthermore, pure hydrogen is produced separately from syngas, bypassing the gas separation process; more importantly, through the rapid cycling of these OCs in different reactors, coking deposition on the catalyst is greatly inhibited [88,89,90,91]. When OCs with high reducibility are applied, methane is fully oxidized to CO2 and H2O (CH4 + 4 OL → CO2 + 2 H2O), which is a strong exothermic reaction, and the released heat can be used for the next step water-splitting reaction, achieving heat self-sufficiency. Another dramatic difference between such a process with CL-SMR is that a third step by air oxidation is normally required to fully regenerate the OCs due to their high oxidation potential (Figure 2B2). This process (three reactors chemical looping M-TRCL) enables hydrogen production with low CO2 emissions, since the high concentration of CO2 can be captured by simple condensation [92].

2.2.1. Supported or Doped Iron Oxides

As for methane-driven three step CLWS, the studied OCs are mainly focused on iron oxides, which holds the virtue of low cost and environmentally friendly features [93,94]. However, developing an efficient method to avoid coke formation, which can degrade the hydrogen purity in the water-splitting step, over these OCs during methane atmosphere remains a great challenge [95]. Ku et al. [96] reported that methane dissociation and coke formation occurred over Fe2O3/Al2O3 OCs due to the generation of the FeAl2O4 phase, since the newly formed phase displayed poor oxygen mobility and rendered notably decreased oxygen capacity of OC for chemical looping reactions. To address this issue, Xiang et al. [97] proposed that loading Fe2O3 to MgAl2O4 support greatly suppressed the solid-phase reaction between iron oxides and the support, thereby restraining the carbon formation from CH4 dissociation. Furthermore, they found that K-promoted Fe2O3/Al2O3 further improved the resistance towards coke deposition by decreasing the reduction activity of oxygen carrier. Li et al. [98] found that a combination of mixed ionic-electronic conductive (MIEC) support of La0.8Sr0.2FeO3 with Fe2O3 improved lattice oxygen diffusion from the bulk to surface, which promoted the elimination of carbon during the methane reduction process.
Instead, when syngas is set as aimed products (methane-driven two-step CLWS), the proper OC needs to bear sufficient reactivity for selective methane conversion. There is usually a seesaw effect in reactivity and selectivity due to the proportional relationship between C–H bond activation ability of CH4 and oxidation ability [99]. For instance, common iron oxides (Fe2O3 and Fe3O4) are reactive towards methane conversion, but CO2 production is preferred over these OCs due to the high oxygen capacity [100,101]. To improve the selectivity of syngas at the CL-SMR process, extensive efforts have been made to break the seesaw effect. Fan et al. [102] proposed that decreasing the particle size of Fe2O3 to 3–5 nm by confining these oxides in mesoporous SBA-15 (Fe2O3@SBA-15) enabled notable enhanced syngas selectivity to above 99%. Corresponding DFT calculations demonstrated that the low coordinated Fe and O atoms of Fe2O3 nanoparticles promoted CH4 activation, cleavage of Fe–O bond, and formation of CO while suppressed the further oxidation to CO2 (Figure 7). Based on DFT + U calculations, they further analyzed the mechanism for CH4 partial oxidation over α-Fe2O3 and claimed that increasing the oxygen vacancy concentration decreased the activation barriers for breaking the C–H bonds of CH4 and Fe–C bonds, which contributed to improved reactivity and syngas selectivity (Figure 8) [103]. Recently, Wang’s team [104] discovered that Y-modified Fe2O3/Al2O3 (Y: Fe = 1.5) could produce a new phase of Y3Fe2Al3O12 garnet after high temperature treatment (1200 °C), which increased the dispersion of active metal Fe, and changed the coordination environment of oxygen species. Lattice oxygen could coordinate with Fe, and also coordinate with Al and Y (Fe-O-Al, Fe-O-Y), which renders increased oxygen vacancy formation energy from 1.4 eV for bear Fe2O3 to 1.997 eV of Y3Fe2Al3O12, which substantially avoids excessive oxidation of CH4 to CO2 while maintaining CO selectivity as high as 98%.

2.2.2. Supported or Doped Cerium Oxides

Cerium-based OCs have been widely studied for CL-SMR in terms of their high oxygen storage capacity and reliable resistance to carbon deposition [105,106,107]. However, pure CeO2 exhibits low reactivity towards methane conversion and is prone to sintering in several cyclic reactions. Therefore, great effort has been applied to improve the reactivity and stability of these OCs by doping foreign cations and constructing composite oxides. Wang et al. [108] discovered that Ni-modified (5 wt%) CeO2-TiO2 enabled 100% conversion of CH4, ca. 16 times greater than that of pure CeO2-TiO2 (5.9%), with 85% syngas selectivity at 900 °C. A mechanism study showed that Ni species accounted for the main active centers for CH4 activation and promoted the reduction of Ce4+ (CeO2-TiO2) to Ce3+ (Ce2Ti2O7) (Figure 9B). The synergy between active Ni species and CeO2-TiO2/Ce2Ti2O7 redox oxides significantly increased the water-splitting performance with hydrogen yield up to 47.0 mL/g (Figure 9). Zhu et al. [87,109,110] prepared a series of Fe-doped cerium oxides (Ce1−xFexO2−δ) for a CL-SMR process and discovered that introduction of iron notably improved the reducibility of OCs, and formation of Ce-Fe-O solid solution (CeFeO3) could enhance the lattice oxygen migration, leading to increased methane conversion. As for the optimized sample (x = 0.5), conversion of methane approached 60% with syngas selectivity exceeding 80% and hydrogen productivity of 83.3 mL/g. Li et al. [111] loaded Ce-Fe-Zr-O oxides (40 wt%) onto inert MgAl-LDO (magnesium–aluminum layered double oxides) and demonstrated that the redox reactivity and stability were greatly improved due to the highly dispersed active components, accelerated reaction between Ce-Fe oxides, and restrained sintering of OCs. Experimental results showed that methane conversion decreased from 58% to 46% in the first 6 cycles and stayed steady in the following 14 cycles with CO selectivity maintained over 80% and hydrogen yield of 1.0 mmol/g. This group further studied the synergy between CeO2 and 3DOM-LaFeO3 during CL-SMR reactions and found that 10 wt% CeO2/LaFeO3 displayed the highest yields of syngas (9.94 mmol/g) and hydrogen (3.38 mmol/g), since the interaction of CeO2-LaFeO3 generated a great number of oxygen vacancies that promoted lattice oxygen diffusion and restricted the carbon deposition [112]. Zhu et al. [113] also found an obvious synergetic effect between CeO2 and BaFe3Al9O19 (BF3). The difference was that a new phase of CeFexAl1−xO3 was generated during chemical looping reactions, which resulted in promoted diffusion of lattice oxygen and increased methane conversion to ca. 90%. Wei et al. [114] investigated the influence of OCs morphology on the redox reactivity and found that CeO2/ZrO2 oxides calcined at low temperatures (450 °C) held stable 3DOM structure with uniform particle size, which effectively enhanced the migration of oxygen species and improved the reactivity for CH4 partial oxidation and water-splitting. After 10 cycles at 800 °C, a hydrogen yield exceeding 38.89 mL/g was maintained.

2.2.3. Perovskites

Perovskite oxides are promising OCs for CL-SMR due to their excellent redox activity [115,116]. Lee et al. [117] discovered that doping Fe in B-site of LaCoO3 improved lattice oxygen mobility, reactivity for partial oxidation of CH4, and adsorption and dissociation of H2O, which enabled high CO selectivity of 92% and H2 purity of 99.3% during the methane reduction and water-splitting steps, respectively. Gong and colleagues [118,119] discovered that substitution of Ce3+ for La3+ of LaFeO3 was capable of tuning lattice oxygen activity via modulating the distortion degree of FeO6 octahedra. When Ce/La ratio reached 1 (La0.5Ce0.5FeO3), the oxygen carrier had the lowest oxygen vacancy formation energy and the highest oxygen mobility, which rendered the best methane-to-syngas and water-splitting performance. Furthermore, the yield of hydrogen (0.6 mmol/g) remained stable during 100 redox cycles (Figure 10), highlighting the outstanding stability of this OC. Recently, Wang et al. [120] proposed that the reactivity of lattice oxygen in LaFe0.8M0.2O3 (M = Al, Fe, Ga, Sc) perovskite was not only influenced by B-site cations, but also greatly affected by the La cations at A-site. It was found that a decline of FeO6 tilting degree by doping smaller cations at B-site of LaFe0.8M0.2O3 was beneficial to weaken the La–O interaction, thereby improving the mobility of lattice oxygen and methane conversion performance. The researchers [121] further showed that doping Sr and Al at A and B sites of LaFeO3, respectively, significantly promoted the deep reduction of Fe4+/3+ cations to Fe0 by forming Fe0@oxides intermediate with core-shell structure. Moreover, the core-shell structure formed in methane atmosphere was able to prevent direct contact between CH4 and Fe0, which promoted syngas production and suppressed coke deposition over Fe0 (Figure 11). Li’s group [122] discovered that doping of Fe in LaMnO3 greatly enhanced the redox performance due to lattice distortion and increased concentration of oxygen vacancy. Among the screened oxides, La0.85MnFe0.15O3 displayed the highest CO selectivity of ~99% and hydrogen yield of 1.76 mmol/g. Similar results could be observed for Ni or Co modified LaMnO3 [123]. Li et al. [124] designed an OCs of Fe2O3@LaxSr1−xFeO3 (LSF) with core-shell structure and found that LSF not only improved the diffusion of lattice oxygen in Fe2O3 but also reduced the direct contact between CH4 and Fe2O3, which resulted in highly selective CH4-to-syngas conversion with selectivity as high as 89%. Zhao et al. [125] found that LaFe1−xCoxO3 (x = 0, 0.2, 0.4, 0.6, 0.8) with a 3DOM structure promoted the mass transfer and provided more active centers for CH4 conversion. Moreover, Co doping increased oxygen mobility, which reduced carbon deposition and improved the purity of hydrogen. This group [126] further investigated the synergistic effect of different metals in double perovskite La1.6Sr0.4FeCoO6, and found that CH4 was prone to be activated on cations with high valence state (Fe5+, Fe4+, and Co3+), while the correspondingly formed Fe2+, Co0 and the oxygen vacancies provided rich active centers for splitting of H2O.

2.3. Chemical Looping Water Gas Shift Process

As a low-cost but important platform chemical, CO can be used as a reductant-like methane to drive the CLWS process and produce high-value hydrogen (Figure 2C) [127,128]. Herein, OCs were reduced by CO (CO + OL → CO2) and subsequently regenerated by water (H2O → OL + H2). This process is generally known as a chemical looping water gas shift (CL-WGS) reaction due to the same overall reaction as that of water–gas shift reaction (CO + H2O → CO2 + H2). Compared with CL-SMR, the application of CO as a reducing atmosphere can avoid the carbon deposition caused by CH4 decomposition while the coke formation from the Boudouard reaction (2CO → C + CO2) is thermodynamically constrained at a temperature above 800 °C, which inhibited the contamination of hydrogen by carbon monoxide (C + H2O → CO + H2). Furthermore, the separation of CO/CO2 from H2O/H2 in chemical looping reactions also prevents the reverse water–gas shift reaction (CO2 + H2 → CO + H2O), leading to increased efficiency for water-splitting. Thermodynamic analysis demonstrated that the steam conversion in CL-WGS reaction reached 95% at 800 °C, which is significantly higher than the traditional WGS process [32]. Based on the ASPEN Plus simulation calculation, Zeng et al. [129] found that the thermal efficiency of H2 production by a chemical looping process, including reduction of OCs by coal and CO (reducer reactor), partial oxidation of OC by steam oxidation to produce hydrogen (oxidizer reactor), and completely recovery of OC by air (combustor), could reach 78% for HHV (higher heating value) efficiency, which was much higher than that of hydrogen production from the traditional coal gasification method.
As for CL-WGS, the suitable OCs should be reactive for CO oxidation and subsequent water-splitting while possessing high redox stability. Fe2O3 has been widely studied due to the features of abundant reserves, favorable thermodynamic properties, and high oxygen-carrying capacity [130]. However, Fe2O3 is prone to agglomerate during several redox cycles, rendering rapid decline of hydrogen productivity. Recent works showed that loading of Fe2O3 on some inert supports, including ZrO2 [131], Al2O3 [132,133], and MgAl2O4 [134], were effective in improving the specific surface area of OCs and slowing down the sintering process. Ma et al. [135] investigated in detail the effects of redox inert supports on the reactivity, redox stability, and hydrogen yield of Fe2O3/support OCs (support = Al2O3, SiO2, MgAl2O4, and ZrO2). It was found that the hydrogen yield of Fe2O3/Al2O3 and Fe2O3/SiO2 gradually decreased with cyclic numbers, since solid phase reaction between Al2O3 or SiO2 and Fe2O3 induced formation of less reactive FeAl2O4 and FeSiO3, respectively. In contrast, when MgAl2O4 was utilized, the solid phase interaction between Fe2O3 and support was greatly inhibited, leading to improved redox stability and hydrogen yield of 54~93 mL/g. Yuzbasi et al. [136] reported that Na-modified Fe2O3/Al2O3 could generate a layered Na-β-Al2O3 phase to restrict the formation of FeAl2O4, which increased the redox stability and rendered a stable hydrogen production of 13.3 mmol/g during 15 cycles. Compared with these inert supports, introduction of MIEC support showed a much-pronounced effect in improving the efficiency of hydrogen production by enhancing migration of lattice oxygen and electron conductivity [137]. Ma et al. [138] studied the performance of AxCe1−xO2−δ (A = Y, Sm and La) supported Fe2O3 for CL-WGS and found that Sm doping (Fe2O3/Ce0.7Sm0.2O1.9) exhibited the best oxygen anion conductivity and oxygen mobility, which facilitated the conversion of CO and subsequent hydrogen production. In contrast, the reactivity of OCs was negatively correlated with La content due to the decreased oxygen capacity arising from formation of LaFeO3 during the reaction. They further discovered that Zr and Sm co-doped Fe2O3/Ce0.8A0.2O2−δ (A = Zr and Sm) displayed better CO conversion and sintering resistance, rending a hydrogen yield of 54.25 mL/g for water-splitting [139].
Iron spinels that bear features of low cost, high thermal stability, and good oxygen mobility are promising OCs for chemical looping reactions [140,141]. Huang et al. [142] found that NiFe2O4/Al2O3 displayed stable H2 yield during 20 cycles of CL-WGS, since Al2O3 support significantly inhibited the sintering of NiFe2O4. Kim et al. [143] also found that Al-modified NiFe2O4 exhibited good performance for water-splitting due to the formation of a spinel solid solution with Al, which could prevent densification of NiFe2O4. For the optimized sample with Al content of 3.3 wt%, maximum hydrogen productivity of 8.2 mmol/g was reached. Cui et al. [144] reported that the active Fe could completely exsolve from mixed Zn-Fe-Al spinel in CO atmosphere and resolved into spinel structure by water oxidation, enabling a hydrogen yield of 2.23 mmol/g with nearly 100% conversion of CO. More importantly, an embedded interface structure was formed for the reduced OC, which effectively limited Fe agglomeration and improved the reactivity and stability of the OCs. Based on DFT+U calculations, Xiao et al. [145] demonstrated that the reactivity of binary spinel AFe2O4 (A = Co, Cu, Ni, Mn) was mainly related to the concentration of oxygen vacancies (reduction level of OCs). Among these OCs, Co-modified OC was the best candidate for CL-WGS due to its suitable oxygen vacancy formation energy (2.70 eV) and lattice oxygen diffusion energy (1.03 eV), which gave an impressive yield of 11.9 mmol/g with average hydrogen production rate of 0.051 mmol/g·s at 650 °C (Figure 12). They further discovered that the oxygen vacancy formation energy and oxygen diffusion energy further decreased to 1.69 eV and 0.83 eV, respectively, after further doping of Cu in CoFe2O4 (Cu0.25Co0.25Fe2.5O4), which promoted the reduction kinetics in CO environment and increased the redox efficiency for hydrogen production. As a result, a hydrogen production as high as ~11.9 mmol/g was achieved at a temperature of 550 °C [146].
Brownmillerite (Ca2Fe2O5) is well known for high oxygen capacity and excellent thermal stability due to the anion-deficient structure with alternating FeO6 octahedral and FeO4 tetrahedral layers [147,148]. Based on the thermodynamic study, Ismail et al. [149] found that the valence state of Fe in Ca2Fe2O5 only displayed Fe0 and Fe3+ without intermediate Fe2+ during the redox reactions, and Fe0 could be directly oxidized to Fe3+ by steam, which was suitable for a water-splitting reaction. Chan et al. [150] discovered that the steam conversion over Ca2Fe2O5 could reach 75%, which was much higher than that of Fe2O3/ZrO2 (62%). Through thermogravimetric experiments, Sun et al. [151] found that Ca2Fe2O5 displayed much faster reaction kinetics than CaFe2O4 during the CO reduction step, rendering a higher reduction degree of 94.0% (85.5% for CaFe2O4). Guo et al. [152] employed DFT calculations to explore the mechanisms of CO adsorption, oxidation, and oxygen anion diffusion on the surface of Ca2Fe2O5 (010). It was discovered that CO preferred to adsorb on FeO2- and FeO-terminated surfaces rather than CaO- and OCa-terminated surfaces, while the chemisorption energy of FeO2-terminated surfaces was considerably greater than that of FeO-terminated surfaces. In addition, it is also proposed that the oxygen anion diffusion at high temperature was the rate-limiting step because the energy barrier of oxygen diffusion was significantly greater than that of the CO oxidation process (Figure 13).

2.4. Syngas Chemical Looping Process

Syngas that can be produced via various routes, e.g., gasification of biomass, coal, and methane reforming, and represents another promising reducing agent for CLWS [153,154,155]. In the syngas-promoted CLWS process (Figure 2D), referred to as syngas chemical looping (SCL), syngas was utilized as the reducing gas instead of CO, while subsequent water-splitting and air oxidation (in some cases) were needed to recover the OCs. The focused OCs explored for this process are mainly iron-based oxides.
Fan et al. [33] investigated the feasibility of Ni, Cu, Cd, Co, Mn, Sn, and Fe oxides for SCL through thermodynamic analysis, and found that Fe2O3 exhibited the best syngas and steam conversion. The Fe2O3 OC displayed stable redox performance during 10 cycles at 600 °C and possessed a low attrition rate of 0.57% during the redox process in entrained flow reactor [156]. When a moving bed reactor was utilized, they found that a syngas conversion of 99.95% with Fe2O3 conversion close to 50% was achieved at 900 °C [157]. Dou et al. [158] showed that the syngas generated by in situ decomposition of glycerol was effective for reduction of OCs. Among the investigated OCs with different Ce/Ni/Fe molar ratios (0/0/1; 10/1/100, 10/3/100, 5/1/100, 5/2/100), OC with Ce/Ni/Fe ratio of 10/3/100 displayed the highest H2 yield of 11.79 mmol/g at 750 °C. Mechanism studies suggested that Ni dopants promoted glycerol conversion while Ce could effectively improve the dispersion of Ni, the sintering resistance, and lattice oxygen mobility, thus substantially boosting the hydrogen yield. Sun et al. [159] investigated the performance of CeO2-modified Fe2O3/Al2O3 in SCL process and found that CeO2 could coordinate with Fe to generate more oxygen vacancies, which promoted the participation of lattice oxygen in the reaction and inhibited the formation of carbon or Fe3C. The best OC with Fe2O3/CeO2/Al2O3 ratio of 65/5/30 OCs achieved the best hydrogen productivity, while the overall performance only decreased by 21% during eighth cycles.
Based on thermodynamic analysis and packed bed experiments, Aston et al. [160] found that the conversion of syngas reached more than 99% during the reduction process of NiFe2O4 and CoFe2O4. Furthermore, the reduced NiFe2O4 (NiFe alloy) and CoFe2O4 (CoFe alloy) was highly reactive towards water-splitting with regeneration of spinel structure, which rendered the NiFe2O4 and CoFe2O4, bearing better redox performance than Fe2O3 (not recovered by water) for SCL process. He et al. [161] investigated the effect of preparation methods, including solid-state method, coprecipitation method, hydrothermal method, and sol-gel method, on the SCL performance of NiFe2O4 nanoparticles. It was found that the OC prepared by sol-gel method displayed the best hydrogen yield and highest lattice oxygen recovery degree due to its smaller particle size and porous structure. Furthermore, further mechanical mixing of NiFe2O4 and inert SiO2 could effectively restrain the sintering of NiFe2O4, rendering outstanding redox stability at 850 °C for 20 cycles. Scheffe et al. [162] prepared m-ZrO2 (surface area of 50 m2/g) supported Co0.85Fe2.15O4 via atomic layer deposition (ALD) method and found that such OC produced much more hydrogen than that over γ-Fe2O3/m-ZrO2 in 7 cycles. Mechanism studies suggested that the high yield of hydrogen was ascribed to the deep reduction of Fe/Co cations into CoFe alloy, exceeding the reduction degree of γ-Fe2O3 (γ-Fe2O3→FeO), which improved the oxygen capacity and water-splitting reactivity, and thus, the hydrogen productivity (Figure 14).

2.5. Photo-Thermochemical Cycle

In 2016, Zhang and coworkers [34] first proposed a strategy of photo-thermochemical cycle (PTC) for CLWS, which explored the photochemical reduction method instead of thermal reduction. This method not only considerably decreased the threshold for abstracting lattice oxygen from OC, slowed down the sintering process, and improved the cyclic stability, but also enabled transformation of solar energy into chemical energy. In a typical photo-thermochemical cycle for water-splitting, the OCs were firstly reduced using ultraviolet and visible light at room temperature with releasing oxygen to generate photo-induced oxygen vacancies, and subsequently recovered by water-splitting via infrared heating to temperature of 500~600 °C with production of hydrogen.
Up to now, the most studied OCs for PTC have been restricted to pure or modified TiO2 photocatalysts. In 2016, TiO2 was used as OC for PTC water-splitting to produce hydrogen, and the average hydrogen production of each cycle reached 0.421 mL/g in five consecutive cyclic reactions [34]. However, TiO2 has a narrow photoresponse range, high recombination rate of photo-induced electron-hole pairs (EHPs), and low reactivity with water. To address these problems, Xu et al. [163] prepared 0.5 wt% Fe/TiO2 as an OC, which displayed much higher hydrogen yield of 33.36 × 10−3 mmol/g than TiO2 (18.79 × 10−3 mmol/g). This is ascribed to the lower recombination rate of EHPs (Figure 15A) and wider photoresponse range (Figure 15B) after doping Fe (III) cations. Further DFT calculations showed that introduction of Fe cations could reduce the oxygen vacancy formation energy, improve H2O adsorption, and enhance the desorption of generated H2, which substantially boosted the efficiency of H2 production (Figure 15C). Docao et al. [164] reported that doping Cu in TiO2 could raise the distortion degree of TiO6 octahedron, which improved the oxygen mobility and promoted water-splitting at 140 °C, with solar energy conversion efficiency reaching 0.63%.
Wu et al. [165] modified TiO2 OCs by doping various transition metals, including Fe, Cu, Co, Ni, and Zn, and discovered that these modified OCs showed higher oxygen vacancy concentration and wider photoresponse range than bare TiO2. Among these OCs, Cu-doped TiO2 displayed the highest light absorption rage and the lowest recombination rate of EHPs, which notably improved the driving force for the release of lattice oxygen and hydrogen production. Further decoration of Cu/TiO2 by Ni could generate more oxygen vacancies and enhance the activity for water-splitting, which exhibited the maximum hydrogen output of 27.01 × 10−3 mmol/g, nearly 36.0 times higher than that of undoped TiO2 (0.75 × 10−3 mmol/g) (Figure 16). The follow-up research showed that 0.1~2.0 wt% of Ni-promoted Cu/TiO2 could anchor nickel oxides (Ni2O3 and NiO) around the surface oxygen vacancy while the low valent state NiO showed more pronounced promotion effect towards water-splitting, which rendered the 0.5 wt% Ni-promoted Cu/TiO2 with highest NiO content, displaying a hydrogen yield (30.6 × 10−3 mmol/g) 40 times higher than that of undoped TiO2 [166].

3. Summary and Outlook

Chemical looping water-splitting is promising for sustainable hydrogen production due to the virtue of decoupling a one-step reaction into two or three spatially separated reactions, which greatly simplifies the gas separation process and avoids the harsh conditions for direct water decomposition. To date, different processes have been developed, including a two-step thermochemical water-splitting (TCWS) cycle, methane chemical looping process, chemical looping water gas shift (CL-WGS) cycle, syngas chemical looping (SCL) process, and photo-thermochemical cycle (PTC), with attempts to reduce the energy penalty and CO2 emissions by altering the method to abstract the lattice oxygen from the OCs, wherein the key lies in the manufacture of suitable OCs. Over the past decades, thousands of metal oxides, such as iron-based metal oxides, zinc-based metal oxides, cerium-based metal oxides, titanium-based metal oxides, and composite metal oxides (perovskite, spinel and garnets, etc.), have been prepared for chemical looping reactions. The performance summary of OCs in different processes of CLWS is summarized in Table 1.
For the two-step TCWS cycle, various OCs, such as iron oxides, zinc oxides, cerium oxides, and perovskite have been widely studied. Among them, cerium oxides have attracted particular attention due to their high structural stability and water-splitting conversion. However, the relatively low reduction degree during the redox cycle, rendering a low hydrogen yield (0.72~7.58 mL/g), greatly hampered its practical applications. Recent work showed that perovskite oxides are promising candidates for two-step TCWS with hydrogen yields of up to 3.13 to 10.71 mL/g, since their redox properties can be facilely modulated by tuning the A/B sites. This gives a clue that constructing composite materials to adjust the redox potential suitable for oxygen desorption and water-splitting should be the key for improving the hydrogen productivity. Furthermore, the sintering problems, thermal shocks, and long single redox cycle time, induced by application of high thermal reduction temperature and low water-splitting temperature, represents another huge obstacle for commercialization of this process. Therefore, introducing support to stabilize the OCs and exploring isothermal redox reactions would be effective in improving the process efficiency for hydrogen production.
Compared to the two-step TCWS cycle, introducing reducing gas, such as methane, carbon monoxide, and syngas, to reduce the oxygen carrier is capable of notably decreasing the reaction temperature to below 1000 °C while enhancing the available oxygen capacity, which significantly decreases the energy consumption, slows down the sintering of OCs, and improve the yield of hydrogen to 13.44~267.63 mL/g. As for methane-driven reduction, valuable syngas with H2/CO ratio of two for Fischer–Tropsch synthesis and methanol production is produced when a suitable OC is selected. Upon OCs with high reducibility applied, the reducing gaseous can be totally combusted with generation of high concentration CO2 (and H2O). All these processes greatly inhibit the side reactions and reduce the burden for gas separation and CO2 caption, rendering improved efficiency and lowered cost for hydrogen production. Among the investigated OCs, iron-based oxides are among the most studied materials due to the virtues of low-cost, environmentally benign features with the high capacity to donate lattice oxygen by varying the valence state of Fe cations. Herein, the relatively low reactivity of iron cations towards conversion of these reducing agents and general trade-off between selectivity and hydrogen yield represent the main challenge for widespread application of these techniques. For CL-SMR with the aim of methane partial oxidation, the reactivity for C–H bond cleavage can be improved by introducing reactive promoters (e.g., Ni and Cu), creating oxygen vacancies, and engineering the morphology of OCs to provide more active centers. To break the trade-off between selectivity and hydrogen productivity, the key lies in a balance between oxygen supply and methane decomposition, which sustainably restrains the coke formation and increases the oxygen capacity for water-splitting. Based on available results, constructing OCs with core-shell structure, composite oxides, and poisoning active Fe0 by forming an alloy or covering with oxides are effective methods to inhibit the carbon deposition. When total combustion of reducing agents (processes of methane-driven three step CLWS, CL-WGS, and SCL) is desired, the reduction degree of OC and hydrogen productivity can be improved by introducing MIEC support, alkali promoters, and some active metals (e.g., Co and Cu), which contribute to promote oxygen mobility and oxygen capacity.
In the case of PTC cycle, the reduction of OCs was realized at ambient temperature with hydrogen production at low temperature (140~600 °C), which greatly suppressed the sintering of OCs. More importantly, the solar energy can be well converted to chemical energy, implementing production of green hydrogen. Up to now, the studied OCs are still limited to bare or modified titanium oxide. Although much progress has been achieved in restraining the recombination rate of EHPs and enhancing the reactivity towards water-splitting, which resulted in improved redox performance, the hydrogen yield (0.0168~0.685 mL/g) is still lower than the above CLWS process, because the photo-driven reduction mainly drives the surface lattice oxygen desorption of OCs, and the amount of desorbed oxygen desorption is limited. In addition to exploring advanced OCs that release more oxygen in the photothermal reduction process, it is anticipated that a combination of suitable reducing gas and photothermal reduction would be promising to significantly enhance the efficiency of hydrogen production.
Over the past 30 years, great effort has been made to develop various OCs by engineering the composition of oxides, including simple metal oxides (cerium oxides, iron oxides, titanium oxides, etc.), composite metal oxides (e.g., perovskite, spinel and garnets), and mixed oxides (e.g., supported OCs, Fe2O3-CeO2, and CeO2-TiO2), which greatly improved the redox performance in CLWS processes [31,48,58,97,104,108,110,125,145,167]. B In addition, modulating the microstructure during the redox reactions also exerts a big influence on the performance due to the modulating of surface active centers, concentration of oxygen vacancies, or metal–oxygen bond strength. For example, Li et al. [124] prepared a kind of Fe2O3@LaxSr1−xFeO3−δ (LSF) OC with core-shell structure and showed that the LSF shell with high methane-to-syngas selectivity could reduce the contact between methane and Fe2O3, which enables a high CO selectivity above 98%, while the Fe2O3 core can donate oxygen to the LSF shell that greatly enhances the oxygen capacity of this OC. Qian et al. [78] found that the crystal phase of CaTi0.5Mn0.5O3−δ OCs could change from an orthogonal phase to a cubic phase during the thermal reduction process, which greatly promoted the desorption of lattice oxygen. Furthermore, the OCs with cubic phase integrate the virtues of high entropy and medium enthalpy, which improves the reactivity for hydrogen production via water-splitting. Wang et al. [121] demonstrated that a specific Fe0@ oxide mediate with core-shell structure was formed when La0.6Sr0.4Fe0.8Al0.2O3−δ was reduced by methane, which notably reduced the direct contact between methane and generated Fe0 species, restrained the coke deposition, and improved the syngas productivity. Cui et al. [144] indicated that the active component (Fe) of ZnFeAlOx OCs reversibly exsolved and dissolved from the interface structure of Zn-Fe-Al-O mixed spinel during a CL-WGS process, which can effectively hinder the migration and agglomeration of iron ions and avoid sintering of materials. These results clearly indicate that adjusting the structure changes during redox reactions represents an effective method to modulate the redox performance for CLWS processes.
At present, selection of OCs for CLWS reactions mainly relies on screening method. This is mainly ascribed to the harsh reaction conditions and dynamic structural evolution during redox reactions, which poses a huge challenge for comprehensively understanding the reaction mechanism and designing advanced OCs for CLWS reactions. Future studies should pay more attention to establish a more precise structure–function relationship with the help of in situ characterization, theoretical calculations, and thermodynamic analysis to provide a theoretical basis and development direction for the design of new efficient long-life OCs. Furthermore, according to the pioneering studies, the research focus of OCs is gradually transferred from simple metal oxides to composite oxides (e.g., perovskite) and mixed oxides due to the feasibility of modulating the redox properties by altering the composition of OCs or synergy between different oxides, which bypasses the shortcomings of single metal oxides, and improves the performance of hydrogen production. Therefore, exploring composite oxides to precisely control the metal–oxygen bond strength and mixed oxides to integrate the advantages of different oxides would be an effective strategy for further improving the redox performance of OCs.

Author Contributions

Conceptualization, writing—original draft preparation, W.C.; writing—review and editing, Y.H. (Yue Hu) and W.X.; supervision, writing—review and editing, C.H.; investigation, H.C.; investigation, J.H.; resources, Y.H. (Yujia Han); supervision, writing—review and editing, Y.Z.; resources, X.M.; supervision, X.W. All authors have read and agreed to the published version of the manuscript.

Funding

National Natural Science Foundation of China (No. 21706254, 21978239, and 22178337), Joint Fund of the Yulin University and the Dalian National Laboratory for Clean Energy (No. 2021012), and the Cyrus Tang Foundation are gratefully acknowledged.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Bockris, J. The origin of ideas on a Hydrogen Economy and its solution to the decay of the environment. Int. J. Hydrogen Energy 2002, 27, 731–740. [Google Scholar] [CrossRef]
  2. Safari, F.; Dincer, I. A review and comparative evaluation of thermochemical water splitting cycles for hydrogen production. Energy Convers. Manag. 2020, 205, 112182. [Google Scholar] [CrossRef]
  3. Alves, H.J.; Bley Junior, C.; Niklevicz, R.R.; Frigo, E.P.; Frigo, M.S.; Coimbra-Araújo, C.H. Overview of hydrogen production technologies from biogas and the applications in fuel cells. Int. J. Hydrogen Energy 2013, 38, 5215–5225. [Google Scholar] [CrossRef]
  4. Dincer, I.; Acar, C. Review and evaluation of hydrogen production methods for better sustainability. Int. J. Hydrogen Energy 2015, 40, 11094–11111. [Google Scholar] [CrossRef]
  5. Mahant, B.; Linga, P.; Kumar, R. Hydrogen Economy and Role of Hythane as a Bridging Solution: A Perspective Review. Energy Fuels 2021, 35, 15424–15454. [Google Scholar] [CrossRef]
  6. Navarro, R.M.; Pena, M.A.; Fierro, J.L. Hydrogen production reactions from carbon feedstocks: Fossil fuels and biomass. Chem. Rev. 2007, 107, 3952–3991. [Google Scholar] [CrossRef]
  7. Go, K.S.; Son, S.R.; Kim, S.D.; Kang, K.S.; Park, C.S. Hydrogen production from two-step steam methane reforming in a fluidized bed reactor. Int. J. Hydrogen Energy 2009, 34, 1301–1309. [Google Scholar] [CrossRef]
  8. Lee, K.B.; Beaver, M.G.; Caram, H.S.; Sircar, S. Novel Thermal-Swing Sorption-Enhanced Reaction Process Concept for Hydrogen Production by Low-Temperature Steam–Methane Reforming. Ind. Eng. Chem. Res. 2007, 46, 5003–5014. [Google Scholar] [CrossRef]
  9. Wang, J.; Sakanishi, K.; Saito, I.; Takarada, T.; Morishita, K. High-Yield Hydrogen Production by Steam Gasification of HyperCoal (Ash-Free Coal Extract) with Potassium Carbonate: Comparison with Raw Coal. Energy Fuels 2005, 19, 2114–2120. [Google Scholar] [CrossRef]
  10. Wang, Z.; Li, L.; Zhang, G. Life cycle greenhouse gas assessment of hydrogen production via chemical looping combustion thermally coupled steam reforming. J. Clean. Prod. 2018, 179, 335–346. [Google Scholar] [CrossRef]
  11. Li, J.; Cheng, W. Comparative life cycle energy consumption, carbon emissions and economic costs of hydrogen production from coke oven gas and coal gasification. Int. J. Hydrogen Energy 2020, 45, 27979–27993. [Google Scholar] [CrossRef]
  12. Zhang, H.; Sun, Z.; Hu, Y.H. Steam reforming of methane: Current states of catalyst design and process upgrading. Renew. Sust. Energy Rev. 2021, 149, 111330. [Google Scholar] [CrossRef]
  13. Siefert, N.S.; Shekhawat, D.; Litster, S.; Berry, D.A. Steam–Coal Gasification Using CaO and KOH for in Situ Carbon and Sulfur Capture. Energy Fuels 2013, 27, 4278–4289. [Google Scholar] [CrossRef]
  14. Kodama, T.; Gokon, N. Thermochemical cycles for high-temperature solar hydrogen production. Chem. Rev. 2007, 107, 4048–4077. [Google Scholar] [CrossRef]
  15. Kogan, A. Direct solar thermal splitting of water and on-site separation of the products-II. Experimental feasibility study. Int. J. Hydrogen Energy 1998, 23, 89–98. [Google Scholar] [CrossRef]
  16. Bilgen, E.; Ducarroir, M.; Foex, M.; Sibieude, F.; Trombe, F. Use of solar energy for direct and two-step water decomposition cycles. Int. J. Hydrogen Energy 1977, 2, 251–257. [Google Scholar] [CrossRef]
  17. Marwat, M.A.; Humayun, M.; Afridi, M.W.; Zhang, H.; Abdul Karim, M.R.; Ashtar, M.; Usman, M.; Waqar, S.; Ullah, H.; Wang, C.; et al. Advanced Catalysts for Photoelectrochemical Water Splitting. ACS Appl. Energy Mater. 2021, 4, 12007–12031. [Google Scholar] [CrossRef]
  18. Zhou, R.; Li, D.; Qu, B.; Sun, X.; Zhang, B.; Zeng, X.C. Rutile TiO2(011)-2 × 1 Reconstructed Surfaces with Optical Absorption over the Visible Light Spectrum. ACS Appl. Mater. Interfaces 2016, 8, 27403–27410. [Google Scholar] [CrossRef] [PubMed]
  19. Anpo, M.; Li, J.; Liu, W.; Qi, W.; Song, F. Hydrogen production technology by electrolysis of water and its application in renewable energy consumption. In Proceedings of the 3rd International Conference on Energy Resources and Sustainable Development (ICERSD 2020), Incheon, Republic of Korea, 13–14 January 2020. [Google Scholar]
  20. Laguna-Bercero, M.A. Recent advances in high temperature electrolysis using solid oxide fuel cells: A review. J. Power Sources 2012, 203, 4–16. [Google Scholar] [CrossRef] [Green Version]
  21. Xue, F.; Su, J.; Li, P.; Zhang, Y. Application of Proton Exchange Membrane Electrolysis of Water Hydrogen Production Technology in Power Plant. In Proceedings of the 3rd International Conference on Air Pollution and Environmental Engineering, Xi’an, China, 28–29 September 2020. [Google Scholar]
  22. Megía, P.J.; Vizcaíno, A.J.; Calles, J.A.; Carrero, A. Hydrogen Production Technologies: From Fossil Fuels toward Renewable Sources. A Mini Review. Energy Fuels 2021, 35, 16403–16415. [Google Scholar] [CrossRef]
  23. Li, D.; Xu, R.; Li, X.; Li, Z.; Zhu, X.; Li, K. Chemical Looping Conversion of Gaseous and Liquid Fuels for Chemical Production: A Review. Energy Fuels 2020, 34, 5381–5413. [Google Scholar] [CrossRef]
  24. Moghtaderi, B. Review of the Recent Chemical Looping Process Developments for Novel Energy and Fuel Applications. Energy Fuels 2011, 26, 15–40. [Google Scholar] [CrossRef]
  25. De Vos, Y.; Jacobs, M.; Van Der Voort, P.; Van Driessche, I.; Snijkers, F.; Verberckmoes, A. Development of Stable Oxygen Carrier Materials for Chemical Looping Processes—A Review. Catalysts 2020, 10, 926. [Google Scholar] [CrossRef]
  26. Funk, J.E.; Reinstrom, R.M. Energy Requirements in Production of Hydrogen from Water. Ind. Eng. Chem. Process 1966, 5, 336–342. [Google Scholar] [CrossRef]
  27. Krenzke, P.T.; Fosheim, J.R.; Davidson, J.H. Solar fuels via chemical-looping reforming. Sol. Energy 2017, 156, 48–72. [Google Scholar] [CrossRef]
  28. Kang, K.-S.; Kim, C.-H.; Bae, K.-K.; Cho, W.-C.; Kim, S.-H.; Park, C.-S. Oxygen-carrier selection and thermal analysis of the chemical-looping process for hydrogen production. Int. J. Hydrogen Energy 2010, 35, 12246–12254. [Google Scholar] [CrossRef]
  29. Kim, Y.; Lim, H.S.; Kim, H.S.; Lee, M.; Lee, J.W.; Kang, D. Carbon dioxide splitting and hydrogen production using a chemical looping concept: A review. J. CO2 Util. 2022, 63, 102139. [Google Scholar] [CrossRef]
  30. Zhao, L.; Dou, B.; Zhang, H.; Wang, Z. Oxygen carriers for chemical-looping water splitting to hydrogen production: A critical review. Carbon Capture Sci. Technol. 2021, 1, 100006. [Google Scholar] [CrossRef]
  31. Chen, Z.; Jiang, Q.; Cheng, F.; Tong, J.; Yang, M.; Jiang, Z.; Li, C. Sr- and Co-doped LaGaO3−δ with high O2 and H2 yields in solar thermochemical water splitting. J. Mater. Chem. A 2019, 7, 6099–6112. [Google Scholar] [CrossRef]
  32. de Leeuwe, C.; Hu, W.; Evans, J.; von Stosch, M.; Metcalfe, I.S. Production of high purity H2 through chemical-looping water–gas shift at reforming temperatures—The importance of non-stoichiometric oxygen carriers. Chem. Eng. J. 2021, 423, 130174. [Google Scholar] [CrossRef]
  33. Gupta, P.; Velazquez-Vargas, L.G.; Fan, L.-S. Syngas Redox (SGR) Process to Produce Hydrogen from Coal Derived Syngas. Energy Fuels 2007, 21, 2900–2908. [Google Scholar] [CrossRef]
  34. Zhang, Y.; Chen, J.; Xu, C.; Zhou, K.; Wang, Z.; Zhou, J.; Cen, K. A novel photo-thermochemical cycle of water-splitting for hydrogen production based on TiO2−x/TiO2. Int. J. Hydrogen Energy 2016, 41, 2215–2221. [Google Scholar] [CrossRef]
  35. Li, F.; Fan, L.-S. Clean coal conversion processes—Progress and challenges. Energy Environ. Sci. 2008, 1, 248–267. [Google Scholar] [CrossRef]
  36. Adanez, J.; Abad, A.; Garcia-Labiano, F.; Gayan, P.; de Diego, L.F. Progress in Chemical-Looping Combustion and Reforming technologies. Prog. Energy Combust. Sci. 2012, 38, 215–282. [Google Scholar] [CrossRef] [Green Version]
  37. Zhu, X.; Imtiaz, Q.; Donat, F.; Müller, C.R.; Li, F. Chemical looping beyond combustion—A perspective. Energy Environ. Sci. 2020, 13, 772–804. [Google Scholar] [CrossRef] [Green Version]
  38. Zeng, L.; Cheng, Z.; Fan, J.A.; Fan, L.-S.; Gong, J. Metal oxide redox chemistry for chemical looping processes. Nat. Rev. Chem. 2018, 2, 349–364. [Google Scholar] [CrossRef]
  39. Qasim, M.; Ayoub, M.; Ghazali, N.A.; Aqsha, A.; Ameen, M. Recent Advances and Development of Various Oxygen Carriers for the Chemical Looping Combustion Process: A Review. Ind. Eng. Chem. Res. 2021, 60, 8621–8641. [Google Scholar] [CrossRef]
  40. Tian, M.; Wang, C.; Han, Y.; Wang, X. Recent Advances of Oxygen Carriers for Chemical Looping Reforming of Methane. ChemCatChem 2021, 13, 1615–1637. [Google Scholar] [CrossRef]
  41. Li, D.; Xu, R.; Gu, Z.; Zhu, X.; Qing, S.; Li, K. Chemical-Looping Conversion of Methane: A Review. Energy Technol. 2019, 8, 1900925. [Google Scholar] [CrossRef]
  42. Voitic, G.; Hacker, V. Recent advancements in chemical looping water splitting for the production of hydrogen. RSC Adv. 2016, 6, 98267–98296. [Google Scholar] [CrossRef] [Green Version]
  43. Luo, M.; Yi, Y.; Wang, S.; Wang, Z.; Du, M.; Pan, J.; Wang, Q. Review of hydrogen production using chemical-looping technology. Renew. Sustain. Energy Rev. 2018, 81, 3186–3214. [Google Scholar] [CrossRef]
  44. Abanades, S. Metal Oxides Applied to Thermochemical Water-Splitting for Hydrogen Production Using Concentrated Solar Energy. ChemEngineering 2019, 3, 63. [Google Scholar] [CrossRef] [Green Version]
  45. Muhich, C.L.; Ehrhart, B.D.; Al-Shankiti, I.; Ward, B.J.; Musgrave, C.B.; Weimer, A.W. A review and perspective of efficient hydrogen generation via solar thermal water splitting. Wiley Interdiscip. Rev. Energy Environ. 2016, 5, 261–287. [Google Scholar] [CrossRef]
  46. Nakamura, T. Hydrogen production from water utilizing solar heat at high temperatures. Sol. Energy 1977, 19, 467–475. [Google Scholar] [CrossRef]
  47. Steinfeld, A.; Sanders, S.; Palumbo, R. Design aspects of solar thermochemical engineering—A case study: Two-step water-splitting cycle using the Fe3O4/FeO redox system. Sol. Energy 1999, 65, 43–53. [Google Scholar] [CrossRef]
  48. Han, S.B.; Kang, T.B.; Joo, O.S.; Jung, K.D. Water splitting for hydrogen production with ferrites. Sol. Energy 2007, 81, 623–628. [Google Scholar] [CrossRef]
  49. Kodama, T.; Kondoh, Y.; Yamamoto, R.; Andou, H.; Satou, N. Thermochemical hydrogen production by a redox system of ZrO2-supported Co(II)-ferrite. Sol. Energy 2005, 78, 623–631. [Google Scholar] [CrossRef]
  50. Kodama, T.; Gokon, N.; Yamamoto, R. Thermochemical two-step water splitting by ZrO2-supported NixFe3−xO4 for solar hydrogen production. Sol. Energy 2008, 82, 73–79. [Google Scholar] [CrossRef]
  51. Ishihara, H.; Kaneko, H.; Yokoyama, T.; Fuse, A.; Hasegawa, N.; Tamaura, Y. Hydrogen Production Through Two-Step Water Splitting Using YSZ (Ni,Fe) System for Solar Hydrogen Production. In Proceedings of the ASME 2005 International Solar Energy Conference, Orlando, FL, USA, 6–12 August 2005. [Google Scholar]
  52. Miller, J.E.; Allendorf, M.D.; Diver, R.B.; Evans, L.R.; Siegel, N.P.; Stuecker, J.N. Metal oxide composites and structures for ultra-high temperature solar thermochemical cycles. J. Mater. Sci. 2008, 43, 4714–4728. [Google Scholar] [CrossRef]
  53. Palumbo, R.; Léde, J.; Boutin, O.; Elorza Ricart, E.; Steinfeld, A.; Möller, S.; Weidenkaff, A.; Fletcher, E.A.; Bielicki, J. The production of Zn from ZnO in a high-temperature solar decomposition quench process—I. The scientific framework for the process. Chem. Eng. Sci. 1998, 53, 2503–2517. [Google Scholar] [CrossRef]
  54. Weidenkaff, A.; Reller, A.W.; Wokaun, A.; Steinfeld, A. Thermogravimetric analysis of the ZnO/Zn water splitting cycle. Thermochim. Acta 2000, 359, 69–75. [Google Scholar] [CrossRef]
  55. Ernst, F.; Steinfeld, A.; Pratsinis, S. Hydrolysis rate of submicron Zn particles for solar H2 synthesis. Int. J. Hydrogen Energy 2009, 34, 1166–1175. [Google Scholar] [CrossRef]
  56. Lv, M.; Zhou, J.; Yang, W.; Cen, K. Thermogravimetric analysis of the hydrolysis of zinc particles. Int. J. Hydrogen Energy 2010, 35, 2617–2621. [Google Scholar] [CrossRef]
  57. Abanades, S.; Flamant, G. Thermochemical hydrogen production from a two-step solar-driven water-splitting cycle based on cerium oxides. Sol. Energy 2006, 80, 1611–1623. [Google Scholar] [CrossRef]
  58. Chueh, W.C.; Falter, C.; Abbott, M.; Scipio, D.; Furler, P.; Haile, S.M.; Steinfeld, A. High-flux solar-driven thermochemical dissociation of CO2 and H2O using nonstoichiometric ceria. Science 2010, 330, 1797–1801. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  59. Costa Oliveira, F.A.; Barreiros, M.A.; Haeussler, A.; Caetano, A.P.F.; Mouquinho, A.I.; Oliveira e Silva, P.M.; Novais, R.M.; Pullar, R.C.; Abanades, S. High performance cork-templated ceria for solar thermochemical hydrogen production via two-step water-splitting cycles. Sustain. Energy Fuels 2020, 4, 3077–3089. [Google Scholar] [CrossRef]
  60. Singh, P.; Hegde, M.S. Ce0.67Cr0.33O2.11: A New Low-Temperature O2 Evolution Material and H2 Generation Catalyst by Thermochemical Splitting of Water. Chem. Mater. 2009, 22, 762–768. [Google Scholar] [CrossRef]
  61. Le Gal, A.; Abanades, S.; Flamant, G. CO2 and H2O Splitting for Thermochemical Production of Solar Fuels Using Nonstoichiometric Ceria and Ceria/Zirconia Solid Solutions. Energy Fuels 2011, 25, 4836–4845. [Google Scholar] [CrossRef]
  62. Chen, Z.; Jiang, Q.; An, H.; Zhang, J.; Hao, S.; Li, X.; Cai, L.; Yu, W.; You, K.; Zhu, X.; et al. Platinum Group Metal Catalyst (RuOx, PtOx, and IrOx)-Decorated Ceria-Zirconia Solid Solution as High Active Oxygen Carriers for Solar Thermochemical CO2 Splitting. ACS Catal. 2022, 12, 7719–7736. [Google Scholar] [CrossRef]
  63. Petkovich, N.D.; Rudisill, S.G.; Venstrom, L.J.; Boman, D.B.; Davidson, J.H.; Stein, A. Control of Heterogeneity in Nanostructured Ce1−xZrxO2 Binary Oxides for Enhanced Thermal Stability and Water Splitting Activity. J. Phys. Chem. C 2011, 115, 21022–21033. [Google Scholar] [CrossRef]
  64. Gokon, N.; Suda, T.; Kodama, T. Oxygen and hydrogen productivities and repeatable reactivity of 30-mol%-Fe-, Co-, Ni-, Mn-doped CeO2−δ for thermochemical two-step water-splitting cycle. Energy 2015, 90, 1280–1289. [Google Scholar] [CrossRef]
  65. Kaneko, H.; Miura, T.; Ishihara, H.; Taku, S.; Yokoyama, T.; Nakajima, H.; Tamaura, Y. Reactive ceramics of CeO2–MOx (M = Mn, Fe, Ni, Cu) for H2 generation by two-step water splitting using concentrated solar thermal energy. Energy 2007, 32, 656–663. [Google Scholar] [CrossRef]
  66. Meng, Q.-L.; Lee, C.-i.; Ishihara, T.; Kaneko, H.; Tamaura, Y. Reactivity of CeO2-based ceramics for solar hydrogen production via a two-step water-splitting cycle with concentrated solar energy. Int. J. Hydrogen Energy 2011, 36, 13435–13441. [Google Scholar] [CrossRef]
  67. Lee, C.-i.; Meng, Q.-L.; Kaneko, H.; Tamaura, Y. Solar Hydrogen Productivity of Ceria–Scandia Solid Solution Using Two-Step Water-Splitting Cycle. J. Sol. Energy Eng. 2013, 135, 011002. [Google Scholar] [CrossRef]
  68. Le Gal, A.; Abanades, S.; Bion, N.; Le Mercier, T.; Harlé, V. Reactivity of Doped Ceria-Based Mixed Oxides for Solar Thermochemical Hydrogen Generation via Two-Step Water-Splitting Cycles. Energy Fuels 2013, 27, 6068–6078. [Google Scholar] [CrossRef]
  69. Le Gal, A.; Abanades, S. Catalytic investigation of ceria-zirconia solid solutions for solar hydrogen production. Int. J. Hydrogen Energy 2011, 36, 4739–4748. [Google Scholar] [CrossRef]
  70. Abanades, S.; Legal, A.; Cordier, A.; Peraudeau, G.; Flamant, G.; Julbe, A. Investigation of reactive cerium-based oxides for H2 production by thermochemical two-step water-splitting. J. Mater. Sci. 2010, 45, 4163–4173. [Google Scholar] [CrossRef]
  71. Le Gal, A.; Abanades, S. Dopant Incorporation in Ceria for Enhanced Water-Splitting Activity during Solar Thermochemical Hydrogen Generation. J. Phys. Chem. C 2012, 116, 13516–13523. [Google Scholar] [CrossRef]
  72. Ruan, C.; Tan, Y.; Li, L.; Wang, J.; Liu, X.; Wang, X. A novel CeO2−xSnO2/Ce2Sn2O7 pyrochlore cycle for enhanced solar thermochemical water splitting. AlChE J. 2017, 63, 3450–3462. [Google Scholar] [CrossRef]
  73. Zhu, X.; Li, K.; Neal, L.; Li, F. Perovskites as Geo-inspired Oxygen Storage Materials for Chemical Looping and Three-Way Catalysis: A Perspective. ACS Catal. 2018, 8, 8213–8236. [Google Scholar] [CrossRef]
  74. Takacs, M.; Hoes, M.; Caduff, M.; Cooper, T.; Scheffe, J.R.; Steinfeld, A. Oxygen nonstoichiometry, defect equilibria, and thermodynamic characterization of LaMnO3 perovskites with Ca/Sr A-site and Al B-site doping. Acta Mater. 2016, 103, 700–710. [Google Scholar] [CrossRef] [Green Version]
  75. McDaniel, A.H.; Miller, E.C.; Arifin, D.; Ambrosini, A.; Coker, E.N.; O’Hayre, R.; Chueh, W.C.; Tong, J. Sr- and Mn-doped LaAlO3−δ for solar thermochemical H2 and CO production. Energy Environ. Sci. 2013, 6, 2424–2428. [Google Scholar] [CrossRef]
  76. Barcellos, D.R.; Sanders, M.D.; Tong, J.; McDaniel, A.H.; O’Hayre, R.P. BaCe0.25Mn0.75O3−δ—A promising perovskite-type oxide for solar thermochemical hydrogen production. Energy Environ. Sci. 2018, 11, 3256–3265. [Google Scholar] [CrossRef]
  77. Sai Gautam, G.; Stechel, E.B.; Carter, E.A. Exploring Ca–Ce–M–O (M = 3d Transition Metal) Oxide Perovskites for Solar Thermochemical Applications. Chem. Mater. 2020, 32, 9964–9982. [Google Scholar] [CrossRef]
  78. Qian, X.; He, J.; Mastronardo, E.; Baldassarri, B.; Yuan, W.; Wolverton, C.; Haile, S.M. Outstanding Properties and Performance of CaTi0.5Mn0.5O3−δ for Solar-Driven Thermochemical Hydrogen Production. Matter 2021, 4, 688–708. [Google Scholar] [CrossRef]
  79. Chuayboon, S.; Abanades, S.; Rodat, S. High-Purity and Clean Syngas and Hydrogen Production From Two-Step CH4 Reforming and H2O Splitting Through Isothermal Ceria Redox Cycle Using Concentrated Sunlight. Front. Energy Res. 2020, 8, 00128. [Google Scholar] [CrossRef]
  80. Wang, L.; Ma, T.; Chang, Z.; Li, H.; Fu, M.; Li, X. Solar fuels production via two-step thermochemical cycle based on Fe3O4/Fe with methane reduction. Sol. Energy 2019, 177, 772–781. [Google Scholar] [CrossRef]
  81. He, F.; Li, F. Perovskite promoted iron oxide for hybrid water-splitting and syngas generation with exceptional conversion. Energy Environ. Sci. 2015, 8, 535–539. [Google Scholar] [CrossRef]
  82. Chuayboon, S.; Abanades, S.; Rodat, S. Syngas production via solar-driven chemical looping methane reforming from redox cycling of ceria porous foam in a volumetric solar reactor. Chem. Eng. J. 2019, 356, 756–770. [Google Scholar] [CrossRef]
  83. Welte, M.; Warren, K.; Scheffe, J.R.; Steinfeld, A. Combined Ceria Reduction and Methane Reforming in a Solar-Driven Particle-Transport Reactor. Ind. Eng. Chem. Res. 2017, 56, 10300–10308. [Google Scholar] [CrossRef] [PubMed]
  84. Saha, D.; Grappe, H.A.; Chakraborty, A.; Orkoulas, G. Postextraction Separation, On-Board Storage, and Catalytic Conversion of Methane in Natural Gas: A Review. Chem. Rev. 2016, 116, 11436–11499. [Google Scholar] [CrossRef] [PubMed]
  85. Olivos-Suarez, A.I.; Szécsényi, À.; Hensen, E.J.M.; Ruiz-Martinez, J.; Pidko, E.A.; Gascon, J. Strategies for the Direct Catalytic Valorization of Methane Using Heterogeneous Catalysis: Challenges and Opportunities. ACS Catal. 2016, 6, 2965–2981. [Google Scholar] [CrossRef]
  86. He, Y.; Zhu, L.; Li, L.; Liu, G. Hydrogen and Power Cogeneration Based on Chemical Looping Combustion: Is It Capable of Reducing Carbon Emissions and the Cost of Production? Energy Fuels 2020, 34, 3501–3512. [Google Scholar] [CrossRef]
  87. Zhu, X.; Wei, Y.; Wang, H.; Li, K. Ce–Fe oxygen carriers for chemical-looping steam methane reforming. Int. J. Hydrogen Energy 2013, 38, 4492–4501. [Google Scholar] [CrossRef]
  88. Zhu, Y.; Liu, W.; Sun, X.; Ma, X.; Kang, Y.; Wang, X.; Wang, J. La-hexaaluminate for synthesis gas generation by Chemical Looping Partial Oxidation of Methane Using CO2 as Sole Oxidant. AlChE J. 2018, 64, 550–563. [Google Scholar] [CrossRef]
  89. Shen, Q.; Huang, F.; Tian, M.; Zhu, Y.; Li, L.; Wang, J.; Wang, X. Effect of Regeneration Period on the Selectivity of Synthesis Gas of Ba-Hexaaluminates in Chemical Looping Partial Oxidation of Methane. ACS Catal. 2018, 9, 722–731. [Google Scholar] [CrossRef]
  90. Krenzke, P.T.; Davidson, J.H. Thermodynamic Analysis of Syngas Production via the Solar Thermochemical Cerium Oxide Redox Cycle with Methane-Driven Reduction. Energy Fuels 2014, 28, 4088–4095. [Google Scholar] [CrossRef]
  91. Wang, X.; Du, X.; Yu, W.; Zhang, J.; Wei, J. Coproduction of Hydrogen and Methanol from Methane by Chemical Looping Reforming. Ind. Eng. Chem. Res. 2019, 58, 10296–10306. [Google Scholar] [CrossRef]
  92. Zhang, F.; Zhu, L.; Wang, Y.; Sun, L. Exergy analysis on the process for three reactors chemical looping hydrogen generation. Int. J. Hydrogen Energy 2020, 45, 24322–24332. [Google Scholar] [CrossRef]
  93. De Vos, Y.; Jacobs, M.; Van Der Voort, P.; Van Driessche, I.; Snijkers, F.; Verberckmoes, A. Sustainable iron-based oxygen carriers for Chemical Looping for Hydrogen Generation. Int. J. Hydrogen Energy 2019, 44, 1374–1391. [Google Scholar] [CrossRef]
  94. Go, K.; Son, S.; Kim, S. Reaction kinetics of reduction and oxidation of metal oxides for hydrogen production. Int. J. Hydrogen Energy 2008, 33, 5986–5995. [Google Scholar] [CrossRef]
  95. Kang, K.-S.; Kim, C.-H.; Cho, W.-C.; Bae, K.-K.; Woo, S.-W.; Park, C.-S. Reduction characteristics of CuFe2O4 and Fe3O4 by methane; CuFe2O4 as an oxidant for two-step thermochemical methane reforming. Int. J. Hydrogen Energy 2008, 33, 4560–4568. [Google Scholar] [CrossRef]
  96. Ku, Y.; Wu, H.-C.; Chiu, P.-C.; Tseng, Y.-H.; Kuo, Y.-L. Methane combustion by moving bed fuel reactor with Fe2O3/Al2O3 oxygen carriers. Appl. Energy 2014, 113, 1909–1915. [Google Scholar] [CrossRef]
  97. Zhu, M.; Chen, S.; Ma, S.; Xiang, W. Carbon formation on iron-based oxygen carriers during CH4 reduction period in Chemical Looping Hydrogen Generation process. Chem. Eng. J. 2017, 325, 322–331. [Google Scholar] [CrossRef]
  98. Galinsky, N.L.; Huang, Y.; Shafiefarhood, A.; Li, F. Iron Oxide with Facilitated O2− Transport for Facile Fuel Oxidation and CO2 Capture in a Chemical Looping Scheme. ACS Sustain. Chem. Eng. 2013, 1, 364–373. [Google Scholar] [CrossRef]
  99. Fung, V.; Polo-Garzon, F.; Wu, Z.; Jiang, D.-e. Exploring perovskites for methane activation from first principles. Catal. Sci. Technol. 2018, 8, 702–709. [Google Scholar] [CrossRef]
  100. Mattisson, T.; Lyngfelt, A. Capture of CO2 using chemical-looping combustion. In Proceedings of the First Biennial Meeting of the Scandinavian-928 Nordic Section of the Combustion Institute, Göteborg, Sweden, 18–20 April 2001. [Google Scholar]
  101. Moghtaderi, B. Hydrogen enrichment of fuels using a novel miniaturised chemical looping steam reformer. Chem. Eng. Res. Des. 2012, 90, 19–25. [Google Scholar] [CrossRef]
  102. Liu, Y.; Qin, L.; Cheng, Z.; Goetze, J.W.; Kong, F.; Fan, J.A.; Fan, L.S. Near 100% CO selectivity in nanoscaled iron-based oxygen carriers for chemical looping methane partial oxidation. Nat. Commun. 2019, 10, 5503. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  103. Cheng, Z.; Qin, L.; Guo, M.; Xu, M.; Fan, J.A.; Fan, L.S. Oxygen vacancy promoted methane partial oxidation over iron oxide oxygen carriers in the chemical looping process. Phys. Chem. Chem. Phys. 2016, 18, 32418–32428. [Google Scholar] [CrossRef] [PubMed]
  104. Kang, Y.; Tian, M.; Huang, C.; Lin, J.; Hou, B.; Pan, X.; Li, L.; Rykov, A.I.; Wang, J.; Wang, X. Improving Syngas Selectivity of Fe2O3/Al2O3 with Yttrium Modification in Chemical Looping Methane Conversion. ACS Catal. 2019, 9, 8373–8382. [Google Scholar] [CrossRef]
  105. Li, D.; Li, K.; Xu, R.; Wang, H.; Tian, D.; Wei, Y.; Zhu, X.; Zeng, C.; Zeng, L. Ce1−xFexO2−δ catalysts for catalytic methane combustion: Role of oxygen vacancy and structural dependence. Catal. Today 2018, 318, 73–85. [Google Scholar] [CrossRef]
  106. Han, Y.; Tian, M.; Wang, C.; Kang, Y.; Kang, L.; Su, Y.; Huang, C.; Zong, T.; Lin, J.; Hou, B.; et al. Highly Active and Anticoke Ni/CeO2 with Ultralow Ni Loading in Chemical Looping Dry Reforming via the Strong Metal–Support Interaction. ACS Sustain. Chem. Eng. 2021, 9, 17276–17288. [Google Scholar] [CrossRef]
  107. Murray, E.P.; Tsai, T.; Barnett, S.A. A direct-methane fuel cell with a ceria-based anode. Nature 1999, 400, 649–651. [Google Scholar] [CrossRef]
  108. Ruan, C.; Huang, Z.-Q.; Lin, J.; Li, L.; Liu, X.; Tian, M.; Huang, C.; Chang, C.-R.; Li, J.; Wang, X. Synergy of the catalytic activation on Ni and the CeO2–TiO2/Ce2Ti2O7 stoichiometric redox cycle for dramatically enhanced solar fuel production. Energy Environ. Sci. 2019, 12, 767–779. [Google Scholar] [CrossRef] [Green Version]
  109. Zhu, X.; Wang, H.; Wei, Y.; Li, K.; Cheng, X. Hydrogen and syngas production from two-step steam reforming of methane over CeO2-Fe2O3 oxygen carrier. J. Rare Earths 2010, 28, 907–913. [Google Scholar] [CrossRef]
  110. Zhu, X.; Li, K.; Wei, Y.; Wang, H.; Sun, L. Chemical-Looping Steam Methane Reforming over a CeO2–Fe2O3 Oxygen Carrier: Evolution of Its Structure and Reducibility. Energy Fuels 2014, 28, 754–760. [Google Scholar] [CrossRef]
  111. Yuan, J.; Zhao, Y.; Xu, H.; Lu, C.; Yang, K.; Zhu, X.; Li, K. Layered Mg-Al spinel supported Ce-Fe-Zr-O oxygen carriers for chemical looping reforming. Chin. J. Chem. Eng. 2020, 28, 2668–2676. [Google Scholar] [CrossRef]
  112. Zheng, Y.; Li, K.; Wang, H.; Tian, D.; Wang, Y.; Zhu, X.; Wei, Y.; Zheng, M.; Luo, Y. Designed oxygen carriers from macroporous LaFeO3 supported CeO2 for chemical-looping reforming of methane. Appl. Catal. B 2017, 202, 51–63. [Google Scholar] [CrossRef]
  113. Cheng, Z.; Zhang, L.; Jin, N.; Zhu, Y.; Chen, L.; Yang, Q.; Yan, M.; Ma, X.; Wang, X. Effect of calcination temperature on the performance of hexaaluminate supported CeO2 for chemical looping dry reforming. Fuel Process. Technol. 2021, 218, 106873. [Google Scholar] [CrossRef]
  114. Zheng, Y.; Wei, Y.; Li, K.; Zhu, X.; Wang, H.; Wang, Y. Chemical-looping steam methane reforming over macroporous CeO2–ZrO2 solid solution: Effect of calcination temperature. Int. J. Hydrogen Energy 2014, 39, 13361–13368. [Google Scholar] [CrossRef]
  115. Rydén, M.; Leion, H.; Mattisson, T.; Lyngfelt, A. Combined oxides as oxygen-carrier material for chemical-looping with oxygen uncoupling. Appl. Energy 2014, 113, 1924–1932. [Google Scholar] [CrossRef] [Green Version]
  116. Long, Y.; Yang, K.; Gu, Z.; Lin, S.; Li, D.; Zhu, X.; Wang, H.; Li, K. Hydrogen generation from water splitting over polyfunctional perovskite oxygen carriers by using coke oven gas as reducing agent. Appl. Catal. B 2022, 301, 120778. [Google Scholar] [CrossRef]
  117. Lee, M.; Lim, H.S.; Kim, Y.; Lee, J.W. Enhancement of highly-concentrated hydrogen productivity in chemical looping steam methane reforming using Fe-substituted LaCoO3. Energy Convers. Manag. 2020, 207, 112507. [Google Scholar] [CrossRef]
  118. Zhang, X.; Su, Y.; Pei, C.; Zhao, Z.-J.; Liu, R.; Gong, J. Chemical looping steam reforming of methane over Ce-doped perovskites. Chem. Eng. Sci. 2020, 223, 115707. [Google Scholar] [CrossRef]
  119. Zhang, X.; Pei, C.; Chang, X.; Chen, S.; Liu, R.; Zhao, Z.J.; Mu, R.; Gong, J. FeO6 Octahedral Distortion Activates Lattice Oxygen in Perovskite Ferrite for Methane Partial Oxidation Coupled with CO2 Splitting. J. Am. Chem. Soc. 2020, 142, 11540–11549. [Google Scholar] [CrossRef] [PubMed]
  120. Xia, X.; Chang, W.; Cheng, S.; Huang, C.; Hu, Y.; Xu, W.; Zhang, L.; Jiang, B.; Sun, Z.; Zhu, Y.; et al. Oxygen Activity Tuning via FeO6 Octahedral Tilting in Perovskite Ferrites for Chemical Looping Dry Reforming of Methane. ACS Catal. 2022, 12, 7326–7335. [Google Scholar] [CrossRef]
  121. Huang, C.; Wu, J.; Chen, Y.-T.; Tian, M.; Rykov, A.I.; Hou, B.; Lin, J.; Chang, C.-R.; Pan, X.; Wang, J.; et al. In situ encapsulation of iron(0) for solar thermochemical syngas production over iron-based perovskite material. Commun. Chem. 2018, 1, 55. [Google Scholar] [CrossRef] [Green Version]
  122. Wang, Y.; Zheng, Y.; Wang, Y.; Wang, H.; Zhu, X.; Wei, Y.; Wang, Y.; Jiang, L.; Yang, Z.; Li, K. Evaluation of Fe substitution in perovskite LaMnO3 for the production of high purity syngas and hydrogen. J. Power Sources 2020, 449, 227505. [Google Scholar] [CrossRef]
  123. Yin, X.; Wang, S.; Wang, B.; Shen, L. Perovskite-type LaMn1−xBxO3+δ (B = Fe, CO and Ni) as oxygen carriers for chemical looping steam methane reforming. Chem. Eng. J. 2021, 422, 128751. [Google Scholar] [CrossRef]
  124. Shafiefarhood, A.; Galinsky, N.; Huang, Y.; Chen, Y.; Li, F. Fe2O3@LaxSr1−xFeO3 Core-Shell Redox Catalyst for Methane Partial Oxidation. ChemCatChem 2014, 6, 790–799. [Google Scholar] [CrossRef]
  125. Zhao, K.; Chen, J.; Li, H.; Zheng, A.; Huang, Z.; Wei, G.; Zhao, Z.; Wang, X. Investigation of the relationship between electronic properties and reactivity of 3DOM LaFe1−xCoxO3 for methane reforming to produce syngas. Int. J. Energy Res. 2019, 43, 7120–7134. [Google Scholar] [CrossRef]
  126. Zhao, K.; Zheng, A.; Li, H.; He, F.; Huang, Z.; Wei, G.; Shen, Y.; Zhao, Z. Exploration of the mechanism of chemical looping steam methane reforming using double perovskite-type oxides La1.6Sr0.4FeCoO6. Appl. Catal. B 2017, 219, 672–682. [Google Scholar] [CrossRef]
  127. Wang, S.; Guan, B.Y.; Lou, X.W.D. Construction of ZnIn2S4-In2O3 Hierarchical Tubular Heterostructures for Efficient CO2 Photoreduction. J. Am. Chem. Soc. 2018, 140, 5037–5040. [Google Scholar] [CrossRef] [PubMed]
  128. Bahmanpour, A.M.; Héroguel, F.; Kılıç, M.; Baranowski, C.J.; Artiglia, L.; Röthlisberger, U.; Luterbacher, J.S.; Kröcher, O. Cu–Al Spinel as a Highly Active and Stable Catalyst for the Reverse Water Gas Shift Reaction. ACS Catal. 2019, 9, 6243–6251. [Google Scholar] [CrossRef]
  129. Zeng, L.; He, F.; Li, F.; Fan, L.-S. Coal-Direct Chemical Looping Gasification for Hydrogen Production: Reactor Modeling and Process Simulation. Energy Fuels 2012, 26, 3680–3690. [Google Scholar] [CrossRef]
  130. Bohn, C.D.; Cleeton, J.P.; Müller, C.R.; Chuang, S.Y.; Scott, S.A.; Dennis, J.S. Stabilizing Iron Oxide Used in Cycles of Reduction and Oxidation for Hydrogen Production. Energy Fuels 2010, 24, 4025–4033. [Google Scholar] [CrossRef]
  131. Liu, W.; Dennis, J.S.; Scott, S.A. The Effect of Addition of ZrO2 to Fe2O3 for Hydrogen Production by Chemical Looping. Ind. Eng. Chem. Res. 2012, 51, 16597–16609. [Google Scholar] [CrossRef]
  132. Kierzkowska, A.M.; Bohn, C.D.; Scott, S.A.; Cleeton, J.P.; Dennis, J.S.; Müller, C.R. Development of Iron Oxide Carriers for Chemical Looping Combustion Using Sol–Gel. Ind. Eng. Chem. Res. 2010, 49, 5383–5391. [Google Scholar] [CrossRef]
  133. Chen, S.; Shi, Q.; Xue, Z.; Sun, X.; Xiang, W. Experimental investigation of chemical-looping hydrogen generation using Al2O3 or TiO2-supported iron oxides in a batch fluidized bed. Int. J. Hydrogen Energy 2011, 36, 8915–8926. [Google Scholar] [CrossRef]
  134. Hafizi, A.; Rahimpour, M. Inhibiting Fe–Al Spinel Formation on a Narrowed Mesopore-Sized MgAl2O4 Support as a Novel Catalyst for H2 Production in Chemical Looping Technology. Catalysts 2018, 8, 27. [Google Scholar] [CrossRef] [Green Version]
  135. Ma, S.; Chen, S.; Soomro, A.; Xiang, W. Effects of supports on hydrogen production and carbon deposition of Fe-based oxygen carriers in chemical looping hydrogen generation. Int. J. Hydrogen Energy 2017, 42, 11006–11016. [Google Scholar] [CrossRef]
  136. Yüzbasi, N.S.; Armutlulu, A.; Huthwelker, T.; Abdala, P.M.; Müller, C.R. Na-β-Al2O3 stabilized Fe2O3 oxygen carriers for chemical looping water splitting: Correlating structure with redox stability. J. Mater. Chem. A 2022, 10, 10692–10700. [Google Scholar] [CrossRef] [PubMed]
  137. Ma, S.; Chen, S.; Soomro, A.; Xiang, W. Effects of CeO2, ZrO2, and Al2O3 Supports on Iron Oxygen Carrier for Chemical Looping Hydrogen Generation. Energy Fuels 2017, 31, 8001–8013. [Google Scholar] [CrossRef]
  138. Ma, S.; Chen, S.; Zhao, Z.; Soomro, A.; Zhu, M.; Hu, J.; Wu, M.; Xiang, W. Enhanced Hydrogen Generation for Fe2O3/CeO2 Oxygen Carrier via Rare-Earth (Y, Sm, and La) Doping in Chemical Looping Process. Energy Fuels 2018, 32, 11362–11374. [Google Scholar] [CrossRef]
  139. Ma, S.; Chen, S.; Ge, H.; Song, T.; Lu, P.; Xiang, W. Synergistic Effects of the Zr and Sm Co-doped Fe2O3/CeO2 Oxygen Carrier for Chemical Looping Hydrogen Generation. Energy Fuels 2020, 34, 10256–10267. [Google Scholar] [CrossRef]
  140. Do, J.Y.; Son, N.; Park, N.-K.; Kwak, B.S.; Baek, J.-I.; Ryu, H.-J.; Kang, M. Reliable oxygen transfer in MgAl2O4 spinel through the reversible formation of oxygen vacancies by Cu2+/Fe3+ anchoring. Appl. Energy 2018, 219, 138–150. [Google Scholar] [CrossRef]
  141. Liu, F.; Wu, F.; Liu, J.; Li, Y.; Yang, Y. Experimental and theoretical insights into the mechanism of spinel CoFe2O4 reduction in CO chemical looping combustion. Fuel 2021, 293, 120473. [Google Scholar] [CrossRef]
  142. Huang, Z.; Deng, Z.; Chen, D.; Wei, G.; He, F.; Zhao, K.; Zheng, A.; Zhao, Z.; Li, H. Exploration of Reaction Mechanisms on Hydrogen Production through Chemical Looping Steam Reforming Using NiFe2O4 Oxygen Carrier. ACS Sustain. Chem. Eng. 2019, 7, 11621–11632. [Google Scholar] [CrossRef]
  143. Kim, Y.; Lim, H.S.; Lee, M.; Kim, M.; Kang, D.; Lee, J.W. Enhanced Morphological Preservation and Redox Activity in Al-Incorporated NiFe2O4 for Chemical Looping Hydrogen Production. ACS Sustain. Chem. Eng. 2021, 9, 14800–14810. [Google Scholar] [CrossRef]
  144. Cui, D.; Li, M.; Qiu, Y.; Ma, L.; Zeng, D.; Xiao, R. Improved hydrogen production with 100% fuel conversion through the redox cycle of ZnFeAlOx oxygen carrier in chemical looping scheme. Chem. Eng. J. 2020, 400, 125769. [Google Scholar] [CrossRef]
  145. Li, M.; Du, Y.; Qiu, Y.; Ma, L.; Cui, D.; Zhang, S.; Tippayawong, N.; Zeng, D.; Xiao, R. The use of ferrites as highly active oxygen storage materials for chemical looping hydrogen production under intermediate temperature. Int. J. Hydrogen Energy 2019, 44, 28638–28648. [Google Scholar] [CrossRef]
  146. Zeng, D.; Qiu, Y.; Li, M.; Cui, D.; Ma, L.; Lv, Y.; Zhang, S.; Xiao, R. Ternary Mixed Spinel Oxides as Oxygen Carriers for Chemical Looping Hydrogen Production Operating at 550 °C. ACS Appl. Mater. Interfaces 2019, 11, 44223–44232. [Google Scholar] [CrossRef]
  147. Hirabayashi, D.; Yoshikawa, T.; Mochizuki, K.; Suzuki, K.; Sakai, Y. Formation of brownmillerite type calcium ferrite (Ca2Fe2O5) and catalytic properties in propylene combustion. Catal. Lett. 2006, 110, 155–160. [Google Scholar] [CrossRef]
  148. Shah, V.; Mohapatra, P.; Fan, L.-S. Thermodynamic and Process Analyses of Syngas Production Using Chemical Looping Reforming Assisted by Flexible Dicalcium Ferrite-Based Oxygen Carrier Regeneration. Energy Fuels 2020, 34, 6490–6500. [Google Scholar] [CrossRef]
  149. Ismail, M.; Liu, W.; Chan, M.S.C.; Dunstan, M.T.; Scott, S.A. Synthesis, Application, and Carbonation Behavior of Ca2Fe2O5 for Chemical Looping H2 Production. Energy Fuels 2016, 30, 6220–6232. [Google Scholar] [CrossRef] [Green Version]
  150. Chan, M.S.C.; Liu, W.; Ismail, M.; Yang, Y.; Scott, S.A.; Dennis, J.S. Improving hydrogen yields, and hydrogen:steam ratio in the chemical looping production of hydrogen using Ca2Fe2O5. Chem. Eng. J. 2016, 296, 406–411. [Google Scholar] [CrossRef] [Green Version]
  151. Sun, Z.; Chen, S.; Hu, J.; Chen, A.; Rony, A.H.; Russell, C.K.; Xiang, W.; Fan, M.; Darby Dyar, M.; Dklute, E.C. Ca2Fe2O5: A promising oxygen carrier for CO/CH4 conversion and almost-pure H2 production with inherent CO2 capture over a two-step chemical looping hydrogen generation process. Appl. Energy 2018, 211, 431–442. [Google Scholar] [CrossRef]
  152. Feng, Y.; Wang, N.; Guo, X.; Zhang, S. Reaction mechanism of Ca2Fe2O5 oxygen carrier with CO in chemical looping hydrogen production. Appl. Surf. Sci. 2020, 534, 147583. [Google Scholar] [CrossRef]
  153. You, S.; Ok, Y.S.; Chen, S.S.; Tsang, D.C.W.; Kwon, E.E.; Lee, J.; Wang, C.H. A critical review on sustainable biochar system through gasification: Energy and environmental applications. Bioresour. Technol. 2017, 246, 242–253. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  154. Hu, Z.; Miao, Z.; Chen, H.; Wu, J.; Wu, W.; Ren, Y.; Jiang, E. Chemical looping gasification of biochar to produce hydrogen-rich syngas using Fe/Ca-based oxygen carrier prepared by coprecipitation. J. Energy Inst. 2021, 94, 157–166. [Google Scholar] [CrossRef]
  155. Bracciale, M.P.; Damizia, M.; De Filippis, P.; de Caprariis, B. Clean Syngas and Hydrogen Co-Production by Gasification and Chemical Looping Hydrogen Process Using MgO-Doped Fe2O3 as Redox Material. Catalysts 2022, 12, 1273. [Google Scholar] [CrossRef]
  156. Li, F.; Kim, H.R.; Sridhar, D.; Wang, F.; Zeng, L.; Chen, J.; Fan, L.S. Syngas Chemical Looping Gasification Process: Oxygen Carrier Particle Selection and Performance. Energy Fuels 2009, 23, 4182–4189. [Google Scholar] [CrossRef]
  157. Li, F.; Zeng, L.; Velazquez-Vargas, L.G.; Yoscovits, Z.; Fan, L.-S. Syngas chemical looping gasification process: Bench-scale studies and reactor simulations. AlChE J. 2010, 56, 2186–2199. [Google Scholar] [CrossRef]
  158. Luo, C.; Dou, B.; Zhang, H.; Liu, D.; Zhao, L.; Chen, H.; Xu, Y. Co-production of hydrogen and syngas from chemical looping water splitting coupled with decomposition of glycerol using Fe-Ce-Ni based oxygen carriers. Energy Convers. Manag. 2021, 238, 114166. [Google Scholar] [CrossRef]
  159. Sun, S.; Zhao, M.; Cai, L.; Zhang, S.; Zeng, D.; Xiao, R. Performance of CeO2-Modified Iron-Based Oxygen Carrier in the Chemical Looping Hydrogen Generation Process. Energy Fuels 2015, 29, 7612–7621. [Google Scholar] [CrossRef]
  160. Aston, V.J.; Evanko, B.W.; Weimer, A.W. Investigation of novel mixed metal ferrites for pure H2 and CO2 production using chemical looping. Int. J. Hydrogen Energy 2013, 38, 9085–9096. [Google Scholar] [CrossRef]
  161. Liu, S.; He, F.; Huang, Z.; Zheng, A.; Feng, Y.; Shen, Y.; Li, H.; Wu, H.; Glarborg, P. Screening of NiFe2O4 Nanoparticles as Oxygen Carrier in Chemical Looping Hydrogen Production. Energy Fuels 2016, 30, 4251–4262. [Google Scholar] [CrossRef]
  162. Scheffe, J.R.; Allendorf, M.D.; Coker, E.N.; Jacobs, B.W.; McDaniel, A.H.; Weimer, A.W. Hydrogen Production via Chemical Looping Redox Cycles Using Atomic Layer Deposition-Synthesized Iron Oxide and Cobalt Ferrites. Chem. Mater. 2011, 23, 2030–2038. [Google Scholar] [CrossRef]
  163. Xu, C.; Zhang, Y.; Chen, J.; Lin, J.; Zhang, X.; Wang, Z.; Zhou, J. Enhanced mechanism of the photo-thermochemical cycle based on effective Fe-doping TiO2 films and DFT calculations. Appl. Catal. B 2017, 204, 324–334. [Google Scholar] [CrossRef]
  164. Docao, S.; Koirala, A.R.; Kim, M.G.; Hwang, I.C.; Song, M.K.; Yoon, K.B. Solar photochemical–thermal water splitting at 140 °C with Cu-loaded TiO2. Energy Environ. Sci. 2017, 10, 628–640. [Google Scholar] [CrossRef]
  165. Wu, Q.; Li, Z.; Zhang, X.; Huang, W.; Ni, M.; Cen, K.; Zhang, Y. Enhanced defect-water hydrogen evolution method for efficient solar utilization: Photo-thermal chemical coupling on oxygen vacancy. Chem. Eng. J. 2021, 408, 127248. [Google Scholar] [CrossRef]
  166. Wu, Q.; Li, Z.; Zhang, X.; Xu, C.; Ni, M.; Cen, K.; Zhang, Y. Oxygen-vacancy-anchoring NiO loading towards efficient hydrogen evolution via photo-thermal coupling reaction. J. Energy Chem. 2021, 61, 77–87. [Google Scholar] [CrossRef]
  167. Zeng, D.; Qiu, Y.; Zhang, S.; Ma, L.; Li, M.; Cui, D.; Zeng, J.; Xiao, R. Synergistic effects of binary oxygen carriers during chemical looping hydrogen production. Int. J. Hydrogen Energy 2019, 44, 21290–21302. [Google Scholar] [CrossRef]
Figure 1. Variation of ΔH, TΔS, and ΔG with the temperature for direct water-splitting at 1 atm. Adapted with permission from Ref. [14], copyright 2007, American Chemical Society.
Figure 1. Variation of ΔH, TΔS, and ΔG with the temperature for direct water-splitting at 1 atm. Adapted with permission from Ref. [14], copyright 2007, American Chemical Society.
Catalysts 13 00279 g001
Figure 2. Chemical looping water-splitting process with various reducing methods: (A) two-step water-splitting thermochemical cycle, (B) methane chemical looping process, (C) chemical looping water gas shift process, (D) syngas chemical looping process, and (E) photo-thermochemical cycle.
Figure 2. Chemical looping water-splitting process with various reducing methods: (A) two-step water-splitting thermochemical cycle, (B) methane chemical looping process, (C) chemical looping water gas shift process, (D) syngas chemical looping process, and (E) photo-thermochemical cycle.
Catalysts 13 00279 g002
Figure 3. Schematic for two-step solar thermochemical water-splitting using Fe3O4/FeO redox system. Adapted with permission from Ref. [47], Copyright 1999, Elsevier.
Figure 3. Schematic for two-step solar thermochemical water-splitting using Fe3O4/FeO redox system. Adapted with permission from Ref. [47], Copyright 1999, Elsevier.
Catalysts 13 00279 g003
Figure 4. (A) Correlation between O2-releasing temperature and the amount of hydrogen produced in the reaction with CeO2–MOx. Reprinted with permission from Ref. [65], copyright 2007, Elsevier. (B) Effect of effective ionic radius and valence state of dopants on the reduction degree of cerium. Reprinted with permission from Ref. [66], copyright 2011, Elsevier.
Figure 4. (A) Correlation between O2-releasing temperature and the amount of hydrogen produced in the reaction with CeO2–MOx. Reprinted with permission from Ref. [65], copyright 2007, Elsevier. (B) Effect of effective ionic radius and valence state of dopants on the reduction degree of cerium. Reprinted with permission from Ref. [66], copyright 2011, Elsevier.
Catalysts 13 00279 g004
Figure 5. (A) Evolution of O2 and H2 for CeO2−xSnO2 (x = 0.05–0.20) during redox reactions. (B) H2 evolution rate during water-splitting reaction at 800 °C for CeO2–0.15SnO2 and pure CeO2. Adapted with permission from Ref. [72], Copyright 2017, Wiley.
Figure 5. (A) Evolution of O2 and H2 for CeO2−xSnO2 (x = 0.05–0.20) during redox reactions. (B) H2 evolution rate during water-splitting reaction at 800 °C for CeO2–0.15SnO2 and pure CeO2. Adapted with permission from Ref. [72], Copyright 2017, Wiley.
Catalysts 13 00279 g005
Figure 6. Oxygen vacancy formation energy of Ca0.5Ce0.5MO3 perovskites derived from CaMO3 (orange bars) and CeMO3 (blue bars) oxides. Red dashed lines show the target scope for solar thermochemical hydrogen production (3.2~4.1 eV). Reprinted with permission from Ref. [77], copyright 2020, American Chemical Society.
Figure 6. Oxygen vacancy formation energy of Ca0.5Ce0.5MO3 perovskites derived from CaMO3 (orange bars) and CeMO3 (blue bars) oxides. Red dashed lines show the target scope for solar thermochemical hydrogen production (3.2~4.1 eV). Reprinted with permission from Ref. [77], copyright 2020, American Chemical Society.
Catalysts 13 00279 g006
Figure 7. (A) Morphological characteristics of Fe2O3@SBA-15 OC. (Aa) HR–TEM images of fresh Fe2O3@SBA-15 and typical Fe2O3 nanoparticles (scale bar of 1 nm); (Ab) HR–TEM images of Fe2O3@SBA-15 after 75 cycles and typical Fe2O3 nanoparticles (scale bar of 5 nm); (B) Temperature programmed reduction profiles of Fe2O3@SBA-15 and unsupported Fe2O3; (C) Calculated results of CH4 adsorption on Fe atop site and O atop site of (Fe2O3)n nanoparticles; (D) Energy evolution of CH4 conversion on Fe40O60 nanoparticle and Fe2O3 (001) surface. Adapted with permission from Ref. [102], copyright 2019, Nature.
Figure 7. (A) Morphological characteristics of Fe2O3@SBA-15 OC. (Aa) HR–TEM images of fresh Fe2O3@SBA-15 and typical Fe2O3 nanoparticles (scale bar of 1 nm); (Ab) HR–TEM images of Fe2O3@SBA-15 after 75 cycles and typical Fe2O3 nanoparticles (scale bar of 5 nm); (B) Temperature programmed reduction profiles of Fe2O3@SBA-15 and unsupported Fe2O3; (C) Calculated results of CH4 adsorption on Fe atop site and O atop site of (Fe2O3)n nanoparticles; (D) Energy evolution of CH4 conversion on Fe40O60 nanoparticle and Fe2O3 (001) surface. Adapted with permission from Ref. [102], copyright 2019, Nature.
Catalysts 13 00279 g007
Figure 8. (A) Mechanism for CH4 partial oxidation over α-Fe2O3 (001) surface with oxygen vacancies. Inset of the reaction loop is the activation barrier (eV) for each elementary step. VO represents the surface oxygen vacancy. (B) The correlation between activation energy of C–H bond and the oxygen vacancy concentration of α-Fe2O3 (001) surface. Reprinted with permission from Ref. [103], copyright 2016, Royal Society of Chemistry.
Figure 8. (A) Mechanism for CH4 partial oxidation over α-Fe2O3 (001) surface with oxygen vacancies. Inset of the reaction loop is the activation barrier (eV) for each elementary step. VO represents the surface oxygen vacancy. (B) The correlation between activation energy of C–H bond and the oxygen vacancy concentration of α-Fe2O3 (001) surface. Reprinted with permission from Ref. [103], copyright 2016, Royal Society of Chemistry.
Catalysts 13 00279 g008
Figure 9. (A) Proposed mechanism for MDR–STCDS and MDR–STWS processes over 5Ni/CeO2–TiO2. (B) Normalized XANES profiles for 5Ni/CeO2–TiO2 after various steps during isothermal MDR–STCDS and MDR–STWS reactions at (Ba) the Ce LIII–edge and (Bb) the Ti K-edge, together with the spectra of reference compounds. Adapted with permission from Ref. [108], copyright 2019, Royal Society of Chemistry.
Figure 9. (A) Proposed mechanism for MDR–STCDS and MDR–STWS processes over 5Ni/CeO2–TiO2. (B) Normalized XANES profiles for 5Ni/CeO2–TiO2 after various steps during isothermal MDR–STCDS and MDR–STWS reactions at (Ba) the Ce LIII–edge and (Bb) the Ti K-edge, together with the spectra of reference compounds. Adapted with permission from Ref. [108], copyright 2019, Royal Society of Chemistry.
Catalysts 13 00279 g009
Figure 10. (A) Distortion degree (Δ) of FeO6 octahedra for La1−xCexFeO3 (x = 0, 0.25 0.5, 0.75, 1) and (B) calculated oxygen vacancy formation energy (ΔEf,vac) of La1−xCexFeO3 (x = 0, 0.5, 1). Adapted with permission from Ref. [119], copyright 2020, American Chemical Society. (C) The oxygen surface exchange coefficient (kchem) and (D) bulk diffusion coefficient (Dchem) of La1−xCexFeO3 (x = 0, 0.5, 1). Redox performance of (E) CH4 partial oxidation step and (F) the H2O-splitting step over La5Ce5FeO3 in CL–SRM at 925 °C. Adapted with permission from Ref. [118], copyright 2020, Elsevier.
Figure 10. (A) Distortion degree (Δ) of FeO6 octahedra for La1−xCexFeO3 (x = 0, 0.25 0.5, 0.75, 1) and (B) calculated oxygen vacancy formation energy (ΔEf,vac) of La1−xCexFeO3 (x = 0, 0.5, 1). Adapted with permission from Ref. [119], copyright 2020, American Chemical Society. (C) The oxygen surface exchange coefficient (kchem) and (D) bulk diffusion coefficient (Dchem) of La1−xCexFeO3 (x = 0, 0.5, 1). Redox performance of (E) CH4 partial oxidation step and (F) the H2O-splitting step over La5Ce5FeO3 in CL–SRM at 925 °C. Adapted with permission from Ref. [118], copyright 2020, Elsevier.
Catalysts 13 00279 g010
Figure 11. Proposed structural evolution of La0.6Sr0.4Fe0.8Al0.2O3−δ OC during chemical looping partial oxidation of methane. White boxes represent the coordination unsaturated Fe cations. Reprinted with permission from Ref. [121], copyright 2018, Nature.
Figure 11. Proposed structural evolution of La0.6Sr0.4Fe0.8Al0.2O3−δ OC during chemical looping partial oxidation of methane. White boxes represent the coordination unsaturated Fe cations. Reprinted with permission from Ref. [121], copyright 2018, Nature.
Catalysts 13 00279 g011
Figure 12. (A) The oxygen vacancy formation energy and (B) diffusion energy of lattice oxygen in CoFe2O4 (100), CuFe2O4 (100), NiFe2O4 (100), MnFe2O4 (100), and Fe2O3 (001) surfaces. O, Cu, Co, Ni, Mn, and Fe atoms are shown with red, light blue, dark blue, light grey, purple, golden spheres, respectively. (C) Profiles of the activation energy and (D) the hydrogen yield for different OCs. Reprinted with permission from Ref. [145], copyright 2019, Elsevier.
Figure 12. (A) The oxygen vacancy formation energy and (B) diffusion energy of lattice oxygen in CoFe2O4 (100), CuFe2O4 (100), NiFe2O4 (100), MnFe2O4 (100), and Fe2O3 (001) surfaces. O, Cu, Co, Ni, Mn, and Fe atoms are shown with red, light blue, dark blue, light grey, purple, golden spheres, respectively. (C) Profiles of the activation energy and (D) the hydrogen yield for different OCs. Reprinted with permission from Ref. [145], copyright 2019, Elsevier.
Catalysts 13 00279 g012
Figure 13. (A) Comparison of CO adsorption energies on FeO2-, CaO-, FeO-, and OCa-terminated Ca2Fe2O5 (0 1 0) surfaces. (B) The pathways and energy profiles for oxygen diffusion in Ca2Fe2O5. Adapted with permission from Ref. [152], copyright 2020, Elsevier.
Figure 13. (A) Comparison of CO adsorption energies on FeO2-, CaO-, FeO-, and OCa-terminated Ca2Fe2O5 (0 1 0) surfaces. (B) The pathways and energy profiles for oxygen diffusion in Ca2Fe2O5. Adapted with permission from Ref. [152], copyright 2020, Elsevier.
Catalysts 13 00279 g013
Figure 14. In situ XRD profiles of OCs reduced by 1 vol % H2/1 vol % CO2 in He at 600 °C. (A) Fe2O3 supported on m-ZrO2 (ALD/20.2% mass loading) and (B) Co0.85Fe2.15O4 supported on m-ZrO2 (ALD/19% mass loading). Chemical reduction for 148 min. Adapted with permission from Ref. [162], copyright 2011, American Chemical Society.
Figure 14. In situ XRD profiles of OCs reduced by 1 vol % H2/1 vol % CO2 in He at 600 °C. (A) Fe2O3 supported on m-ZrO2 (ALD/20.2% mass loading) and (B) Co0.85Fe2.15O4 supported on m-ZrO2 (ALD/19% mass loading). Chemical reduction for 148 min. Adapted with permission from Ref. [162], copyright 2011, American Chemical Society.
Catalysts 13 00279 g014
Figure 15. (A) PL profiles for TiO2 and 0.5 wt% Fe-doped TiO2. (B) UV–vis DRS profiles for TiO2 and 0.5 wt% Fe-doped TiO2. Inset shows the optical energy band gap. (C) The reaction pathway of H2O splitting during PTC process. Reprinted with permission from Ref. [163], copyright 2017, Elsevier.
Figure 15. (A) PL profiles for TiO2 and 0.5 wt% Fe-doped TiO2. (B) UV–vis DRS profiles for TiO2 and 0.5 wt% Fe-doped TiO2. Inset shows the optical energy band gap. (C) The reaction pathway of H2O splitting during PTC process. Reprinted with permission from Ref. [163], copyright 2017, Elsevier.
Catalysts 13 00279 g015
Figure 16. (A) Schematic of the reaction mechanism for photochemical and thermochemical processes. (B) The average H2 yield of different catalysts. TR represented “thermochemical reaction”. Reprinted with permission from Ref. [165], copyright 2021, Elsevier.
Figure 16. (A) Schematic of the reaction mechanism for photochemical and thermochemical processes. (B) The average H2 yield of different catalysts. TR represented “thermochemical reaction”. Reprinted with permission from Ref. [165], copyright 2021, Elsevier.
Catalysts 13 00279 g016
Table 1. Performance summary of OCs in different processes in CLWS.
Table 1. Performance summary of OCs in different processes in CLWS.
Hydrogen Production ProcessesOCs TypeCyclic ReactionOCsReaction Temperature (°C; Red/Ox)H2 Yield (mmol/g)Ref.
Two-step TCWS cycleFerrites
(MxFe3−xO4)
MxFe3−xO4 → xMO + (3−x)FeO + 0.5O2;
xMO + (3−x) FeO + H2O → MxFe3−xO4 + H2
NiFe2O41200/8000.0197 *[48]
Co0.67Fe2.33O4/ZrO21300~1400/10000.357~0.535[49]
NiFe2O41400/1000~0.446[50]
NiFe2O4/m-ZrO21400/10000.535~0.714[50]
ZnO2ZnO → Zn + O2;
Zn + H2O → ZnO + H2
ZnO--
Cerium oxide
(CexM1−xO2)
CexM1−xO2 → CexM1−xO2−δ + 0.5δ O2;
CexM1−xO2−δ + H2O → δH2O → CexM1−xO2 + δH2
CeO21350/8000.032[31]
CeO21500/5000.142 *[66]
Ce0.9Hf0.1O21500/5000.2 *[66]
Ce0.9Sc0.1O1.951500/5000.181 *[66]
Ce0.75Zr0.25O21400/10500.297[69]
1% Gd/CeO2-ZrO21400/10500.338 *[71]
CeO2-0.15SnO21400/8000.321 *[72]
CeO21350/8500.05[76]
CeO21500/10000.22[76]
ABO3ABO3 ↔ ABO3−δ + 0.5δO2;
ABO3−δ + δH2O → ABO3 + δH2
LaGa0.4Co0.6O3−δ1350/8000.478[31]
Sr0.4La0.6Mn0.6Al0.4O31350/10000.307[75]
Sr0.4La0.6Mn0.6Al0.4O3−δ1350/8500.194[76]
BaCe0.25Mn0.75O3−δ1350/8500.140[76]
CaTi0.5Mn0.5O3−δ1350/11500.446[78]
Methane chemical looping processThree reactors chemical looping (M-TRCL)Iron oxides
(Fe2O3)
Fe2O3 + CH4 → FeO + 3CO2 + 2H2O;
3FeO + H2O →Fe3O4 + H2;
2Fe3O4 + 0.5O2 → 3Fe2O3
K promoted-Fe2O3/Al2O3850/850-[97]
Fe2O3/La0.8Sr0.2FeO3900/900-[98]
CL-SMRCerium oxide
(CexM1−xO2)
CexM1−xO2 + δCH4 → CexM1−xO2−δ + δCO + 2δH2;
CexM1−xO2−δ + δH2O → CexM1−xO2 + δH2
Ce0.5Fe0.5O2−δ850/8503.71[87]
5Ni/CeO2-TiO2900/9002.098 *[108]
CeO2-Fe2O3 (Ce/Fe = 7/3)850/7001.116[109]
40 wt% Ce-Fe-Zr-O/MgAl-LDO850/8501.0[111]
CeO2-LaFeO3800/8003.38[112]
CeO2/ZrO2800/8001.736[114]
ABO3ABO3 + δCH4 → ABO3−δ + δCO + 2δH2;
ABO3−δ + δH2O → ABO3 + δH2
LaCo0.6Fe0.4O3700/7002.22[117]
La0.5Ce0.5FeO3925/9250.6[118]
La0.85MnFe0.15O3850/8501.76[122]
CL-WGSIron oxides
(Fe2O3)
Fe2O3 + 3CO → 2Fe + 3CO2;
3Fe + 4H2O → Fe3O4 + 4H2;
2Fe3O4 + 0.5O2 → 3Fe2O3
Fe2O3/SiO2900/900~0.745 *[135]
Fe2O3/Al2O3900/900~0.982 *[135]
Fe2O3/YSZ900/900~2.232 *[135]
Fe2O3/ZrO2900/900~2.977 *[135]
Fe2O3/MgAl2O4900/9002.41~4.152 *[135]
Na-modified Fe2O3/Al2O3800/800~13.3[136]
Fe2O3/ZrO2850/850~7.143 *[137]
Fe2O3/CeO2850/850~5.80 *[137]
Fe2O3/Ce0.6Sm0.15Zr0.25O1.925850/8502.422 *[139]
Iron spinels
(MxFe3−xO4)
MxFe3−xO4 + (2−x)H2 + CO → xMO + (3−x)Fe + CO2 + (3−x)H2O;
Fe + H2O → Fe3O4 + H2;
xMO + (3−x)/3Fe3O4 + 0.5xO2 → MxFe3−xO4
3.3 wt% Al-modified NiFe2O4800/8008.2[143]
ZnFeAlOx850/8502.23[144]
CoFe2O4650/65011.9[145]
Cu0.25Co0.25Fe2.5O4550/550~11.9[146]
Brownmillerite (Ca2Fe2O5)Ca2Fe2O5 + 3CO → 2CaO + 2Fe + 3CO2;
2CaO + 2Fe + 3H2O → Ca2Fe2O5 + 3H2
Ca2Fe2O5900/900-[149]
SCL processIron oxides
(Fe2O3)
Fe2O3 + 2H2 + CO → 2Fe + CO2+ 2H2O
3Fe + 4H2O → Fe3O4 + 4H2;
2Fe3O4 + 0.5O2 → 3Fe2O3
Fe-Ce-Ni based OCs (Ce/Ni/Fe ratio of 10/3/100)750/75011.79[158]
Fe2O3600/6001.14[162]
Iron spinels
(MxFe3−xO4)
MxFe3−xO4 + CO → xMO + (3−x)Fe + CO2;
xMO + (3−x)FeO + H2O → MxFe3−xO4 + H2;
Co0.85Fe2.15O4/m-ZrO2600/6003.944[162]
PTCTiO22TiO2 → Ti2O3 + 0.5O2;
Ti2O3 + H2O → 2TiO2 + H2
(hv + TiO2 → e + h+;
e + Ti4+ → Ti3+;
h+ + O2 → O2 + Vo;
Vo + H2O → Vo(H2O)
Ti3+-Vo (H2O)-Ti3+ → Ti4+-O-Ti4+ + H2)
TiO2room
temperature/500-600
0.0188 *[34]
0.5 wt% Fe/TiO2room
temperature/600
0.0334[163]
TiO2room
temperature/350
0.75 × 10−3[165]
Fe/TiO2room
temperature/350
9.73 × 10−3[165]
Cu/TiO2room
temperature/350
7.13 × 10−3[165]
Ni-promoted Cu/TiO2room
temperature/350
0.027[165]
0.5% Ni-promoted Cu/TiO2room
temperature/350
0.0306[166]
* The unit of raw hydrogen production data is “mL/g”, which can be converted into “mmol/g” by the formula PV = nRT with hypothesis of STP condition.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Chang, W.; Hu, Y.; Xu, W.; Huang, C.; Chen, H.; He, J.; Han, Y.; Zhu, Y.; Ma, X.; Wang, X. Recent Advances of Oxygen Carriers for Hydrogen Production via Chemical Looping Water-Splitting. Catalysts 2023, 13, 279. https://doi.org/10.3390/catal13020279

AMA Style

Chang W, Hu Y, Xu W, Huang C, Chen H, He J, Han Y, Zhu Y, Ma X, Wang X. Recent Advances of Oxygen Carriers for Hydrogen Production via Chemical Looping Water-Splitting. Catalysts. 2023; 13(2):279. https://doi.org/10.3390/catal13020279

Chicago/Turabian Style

Chang, Wenxi, Yue Hu, Weibin Xu, Chuande Huang, Haonan Chen, Jiahui He, Yujia Han, Yanyan Zhu, Xiaoxun Ma, and Xiaodong Wang. 2023. "Recent Advances of Oxygen Carriers for Hydrogen Production via Chemical Looping Water-Splitting" Catalysts 13, no. 2: 279. https://doi.org/10.3390/catal13020279

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop