Next Article in Journal
Dependence of the Fluidizing Condition on Operating Parameters for Sorption-Enhanced Methanol Synthesis Catalyst and Adsorbent
Previous Article in Journal
Copper- and Manganese-Based Bimetallic Layered Double Hydroxides for Catalytic Reduction of Methylene Blue
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

(La1−xCax)MnO3−δ (x = 0, 0.2, 0.3, 0.4) Perovskites as Redox Catalysts in Chemical Looping Hydrogen Production Process: The Relation between Defect Chemistry and Redox Performance

by
Moschos Moschos
1,
Antigoni Evdou
2 and
Vassilios Zaspalis
1,2,*
1
Department of Chemical Engineering, Aristotle University of Thessaloniki, University Campus, 54124 Thessaloniki, Greece
2
Chemical Process Engineering Research Institute, Center for Research and Technology-Hellas, Thermi, 57001 Thessaloniki, Greece
*
Author to whom correspondence should be addressed.
Catalysts 2024, 14(7), 431; https://doi.org/10.3390/catal14070431
Submission received: 4 June 2024 / Revised: 2 July 2024 / Accepted: 3 July 2024 / Published: 6 July 2024
(This article belongs to the Section Nanostructured Catalysts)

Abstract

:
The interaction between point defects in (La1−xCax)MnO3−δ (x = 0, 0.2, 0.3, 0.4) perovskites and their redox catalytic properties in a three-reactor chemical looping hydrogen production process is investigated. During the reduction step with CH4, the behavior of the materials is extrinsically determined and strongly depends on the Ca content. At small oxygen deficiencies, CH4 becomes oxidized to CO2. As the deficiency increases, partial oxidation to CO and H2 at a molar ratio of approximately 2 is favored. During the water-splitting step, the dependency on the Ca content is much weaker since it is intrinsically determined by the Mn2+→Mn3+ oxidation with simultaneous annihilation of oxygen vacancies that are not required to compensate for the extra negative charge of the Ca dopant. Hydrogen productivities in the order of 13 cm3 (STP) H2/g solid could be achieved during the water-splitting step at 1000 °C. The materials exhibited reproducible catalytic behavior during 10 cycles of the complete three-step process and were found to retain their perovskite structure.

Graphical Abstract

1. Introduction

Τhere is no doubt that one of the major concerns of humanity is related to the continuously rising emissions of CO2 originating from fossil resources [1]. These anthropogenic emissions are considered responsible for the global temperature rise and consequent climate changes [2]. On the other hand, since fossil fuels will most likely remain the primary energy source for many years, an efficient reduction of CO2 emissions could be accomplished by the development of fuel combustion technologies with integrated CO2 capture or storage. Based on these arguments, Chemical Looping Combustion (CLC) technologies have been developed and extensively investigated [3,4]. The basic idea of this technology is the utilization of a solid oxygen carrier (OC), which is reduced by supplying oxygen for the combustion of fuel in the fuel reactor (FR). This omits the utilization of air as an oxygen supply and thus enables the fuel conversion to a pure stream of CO2 and H2O that can be easily separated by condensation. The oxygen-deficient OC is then transferred to the air reactor (AR), where it is reoxidized by air to its initial state, and the loop cycle is repeated (Figure 1a).
Quite soon after, the basic chemical looping concept was extended in order to cover other industrial needs besides energy production [5,6]. Chemical Looping Reforming (CLR) is a process very similar to CLC (Figure 1a), with the main difference being that the fuel (which is mainly CH4) is partially oxidized to CO and H2 [7,8]. It is attractive for CO2 capture and H2 generation through the reforming process, and when combined with high efficiencies and low H2 production costs, it may provide a good alternative to traditional H2 production technologies. Other chemical looping variations include Syngas Chemical Looping (SCL), which uses syngas (CO+H2) as a fuel to produce H2 and CO2 [9,10], or Coal Direct Chemical Looping (CDCL), which uses coal as a solid fuel in a three-reactor configuration [11].
An interesting variant of the chemical looping concept is the so-called chemical looping hydrogen (CLH) or Chemical Looping Water-Splitting Process (CLWS) (Figure 1b) [12,13]. The fuel (usually CH4) is oxidized in the fuel reactor with the OC, which becomes reduced. The reduced and oxygen-deficient OC is then transferred to the steam reactor, where it is partially oxidized through the water-splitting reaction. Finally, the partially oxidized OC is transferred to the third air reactor, where it is fully oxidized to its initial state, and the loop closes. An advantage of the CLH process is that, next to the hydrogen produced in the FR, very pure H2 is produced at the exit of the steam reactor by simply cooling and condensing the H2/H2O mixture. An important technological challenge is the circulation of solids through the three reactors.
The selection of the proper OC is one of the most important technological issues for all chemical looping technologies and plays a decisive role in the techno-economical evaluation of the process. High oxygen exchange capacity, favorable oxidation or reduction thermodynamics or structural stability are among the requirements. Extensive investigations have been performed to evaluate a large number of OCs for chemical looping processes. Summarized results have been published in several reviews [14,15].
Concerning the CLH process, the research work is almost exclusively concerned with iron oxide-based OCs [12,13,16]. The reduced iron oxide (FeO) in the FR is partially oxidized (Fe3O4) in the SR and totally oxidized (Fe2O3) in the AR. In order to improve the redox performance or the mechanical properties, several additives are added, such as ZrO2 [17], CeO2 [18], CaO [19] and Al2O3 [20].
Perovskites comprise an interesting family of materials known for their high thermal stability and sintering resistance or their high oxygen mobility and ability to accommodate relatively large deviations from stoichiometry. As redox catalysts, they have been reported to have promising properties in CLR applications since they facilitate the partial oxidation of methane to CO and H2 [21,22]. A large number of perovskite materials has been investigated, including La1−xSrxMO3 (M = Mn, Ni) [23,24], La1−xSrxFeO3 [25], LaFe1−xNixO3 [26], LaFe1−xCoxO3 [27] and La1−xCaxMnO3 [28].
The investigations of perovskites for the three-reactor CLH process are rather limited. In certain cases, perovskites have been employed to complement iron oxide-based catalysts in order to improve the redox performance or lifetime [29,30,31]. LaFeO3 perovskites showed quite promising results that could be enhanced with partial substitution of Fe with Co (in order to increase the V O concentration) in the modified compositions LaFe1−xCOxFeO3. An optimum performance could be achieved at x = 0.3 [27,32]. La0.8Ca0.2MO3±δ (M = Co, Ni, Fe, Cu) perovskites have been investigated for the two-step thermochemical water-splitting reaction; the material reduction occurred only upon equilibration under a nitrogen atmosphere (PO2~10−5 atm). The pure hydrogen produced during the water-splitting step approaches 5 cm3 (STP)/g solid with OC reduction and oxidation temperatures of 800 °C [33,34].
In this article, we report on the performance of (La1−xCax)MnO3−δ perovskites as OCs in the three-reactor CLH process. Τhese materials are considered of interest since, due to their high melting point, ionic and electronic conductivity or mechanical stability under oxidizing or reducing conditions, they have been reported to be the preferred materials for several applications, such as solid oxide fuel cells or mixed conducting membranes, [35,36] and are being investigated for various catalytic processes [37,38,39]. In addition, their defect chemistry is well known [40] and shows good stability in chemical looping processes [24]. Attention has been paid to the connection between the properties of interest, such as the products of the various oxidation-reduction reactions, and the nonstoichiometry and defect structure of these compounds, which is described based on well-accepted models published in the literature. These models are mainly constructed based on macroscopic electrical behavior evaluation, although currently, they can be verified using more powerful physical or chemical techniques [41]. Although perovskite materials are well known to exhibit intrinsic or extrinsic point defect concentrations that depend strongly on oxidative or reductive environments and affect their catalytic properties [42], attempts toward understanding the interaction between redox performance and defect chemistry are quite limited in the chemical looping literature.

2. Results

The XRD patterns of the as-prepared samples are shown in Figure 2. From the overall spectrum (shown in the left part of the figure), it is concluded that all samples have the perovskite structure. Detailed examination of the patterns revealed the presence of traces of phases identified as La2O3 or La2CaOx. Due to the low concentrations, these are considered of no importance for the general behavior of the materials in the CLH process. From the detailed spectrum of the peak at 2θ between 32 and 33° (shown in the right part of the figure), it is indicated the substitution of Ca for La transforms the rhombohedral structure to orthorhombic and decreases the unit cell size. This result is in agreement with the results reported and explained in the literature [28,43].
In Figure 3, the evolution of the oxygen stoichiometric parameter δ is shown as the material passes from the reduction step to the water-splitting step and finally to the total oxidation step.
In the reduction step in the fuel reactor, the deficiency initially increases at almost the same rate per pulse while gradually flattening toward some constant value. The pulse number at which deviation from the initial linearity and flattening start to occur increases with increasing Ca content. The same also holds for the deficiency level toward the end of the reduction step. When H2O is subsequently fed to the reactor, water splitting occurs, and the oxygen deficiency starts to decrease toward a constant value. The reduced materials become partly reoxidized. The deficiency difference between the start and the end of the water-splitting test does not seem to strongly depend on the Ca content, at least not so strongly as the deficiency evolution during the reduction step. The deficiency order after the water-splitting step is the same as that at the end of the reduction step. Finally, during the final oxidation, the deficiency further decreases down to a constant value of δ ~ 0.05, which is independent of the Ca content and differs slightly from the deficiency of the initial “as prepared” materials (δ = δref ≈ 0). In the insert of Figure 3, the deficiency evolution is shown for identical experiments without the intermediate water-splitting test. All reduced samples are fully oxidized down to almost their initial state.
In Figure 4, the CH4 conversion per pulse X C H 4 and the cumulative CH4 consumption are shown as a function of the pulse number during the reduction step.
Initially, almost complete CH4 conversion per pulse is achieved, which decreases as the pulse number increases. The higher the Ca content, the higher the CH4 conversion per pulse. Toward the end of the reduction step, in all cases, a steady-state conversion of approximately 20% is achieved, independent of the Ca content. Moreover, the higher the Ca content, the higher the cumulative CH4 consumption. In agreement with the identical steady-state conversions, all cumulative lines end in parallel linear regions that correspond to a constant CH4 consumption of ~8 μmoles CH4 per g solid per pulse.
The cumulative distributions of products CO, CO2, H2 and H2O during the reduction step are shown in Figure 5a,b.
Initially, at small oxygen deficiencies, CO2 and H2O are mainly produced. Subsequently, as the oxygen deficiency increases (with increasing pulse number), the production of CO2 and H2O is almost entirely substituted by the production of CO and H2. The higher the Ca content, the higher the deficiency at which CO and H2 appear and finally dominate the reaction products. At the end of the reduction step, an almost constant production of 5.5–6 μmoles CO per g solid per pulse and 11–12 μmoles H2 per g solid per pulse are produced, independently of the Ca content of the materials. The H2 to CO molar ratio is very close to 2. By comparing Figure 3, Figure 4 and Figure 5, it is observed that toward the end of the 25 pulses of the reduction step, a situation is achieved, in all cases, with a constant CH4 conversion of ~20% toward the production of CO and H2, without significantly affecting the oxygen deficiency of the solids. An explanation will be provided in a subsequent paragraph.
In Figure 6, the cumulative H2 production as a function of the pulse number is shown only during the water-splitting step.
The cumulative H2 production increases as the Ca content increases from 147 μmoles H2/g solid (~3.3 cm3 (STP)/g solid) for the LaMnO3 material up to 255 μmoles H2/g solid (~5.7 cm3 (STP)/g solid) for the (La0.6Ca0.4)MnO3 material.
In Figure 7, the effect of temperature on the evolution of the oxygen deficiency during the reduction and the water-splitting step is shown for the (La1−xCax)MnO3 materials with x = 0.3 and 0.4. Similar results are also achieved by the other tested compositions. Except for the standard temperature of 900 °C, CLH experiments are also performed at 800 °C and 1000 °C. As indicated in Figure 7, the temperature increase does not only cause an increase in the oxygen deficiency evolution during the reduction step but also increases the deficiency reduction (i.e., the H2 production) during the water-splitting step. This is a strong indication that the reactions involved in these two steps are endothermic. It is worthwhile to notice that at 1000 °C, the oxygen deficiency evolution did not reach a steady state, and apparently, in this case, more than 25 pulses were necessary.
Corresponding CH4 conversions and cumulative H2 productions are shown in Figure 8. The increased CH4 conversions with increased reduction temperature are also associated with an increase in the almost constant conversion toward the end of the reduction step. The cumulative H2 production, either during the reduction step or during the water-splitting step, is also increased with increased process temperature. The total H2 production of the (La0.7Ca0.3)MnO3 material is 927 μmoles H2/g solid (~21 cm3 (STP)/g solid), from which 352 μmoles H2/g solid (~8 cm3 (STP)/g solid) are produced as syngas during the reduction step and 575 μmoles H2/g solid (~13 cm3 (STP)/g solid) as pure H2 during the water-splitting step.
The stability and performance reproducibility of the (La1−xCax)MnO3 materials have been tested for 10 successive reduction–partial oxidation–total oxidation cycles. The results are shown in Figure 9 for material (La0.6Ca0.4)MnO3. Similar results are also obtained with the other tested compositions. As shown, the materials exhibit a stable and reproducible oxidation-reduction behavior after 10 cycles. The oxygen deficiency of the material after each process step is considered reproducible with cycle repetition.
In Figure 10, XRD results are shown for samples (La0.7Ca0.3)MnO3 and (La0.6Ca0.4)MnO3 after each step of the CLH process. For sample (La0.6Ca0.4)MnO3, results are shown after seven reduction–total oxidation cycles and after ten reduction–partial oxidation–total oxidation cycles. Similar results are also achieved with all compositions. In all cases, the main perovskite structure is preserved, and no decomposition into low- or high-valence oxides is observed. The appearance of some minor secondary phases is not considered important for the general material behavior. The reduction step clearly leads to larger cell dimensions since lattice oxygen is consumed for the oxidation of CH4, and oxygen vacancies are created. Both partial oxidation with H2O and further total oxidation with O2 restore oxygen deficiency and decrease the orthorhombic unit cell size. However, it never returns to its initial size. This result is in qualitative agreement with the fact that the oxygen deficiency after total oxidation does not return to the reference deficiency of the fresh samples (i.e., δ = δref ≈ 0) and indicates the existence of some non-reversible phenomena that affect the structure, which cannot be explained at the moment. Nevertheless, the structural and lattice size similarity after 7 CH4/O2 cycles and 10 CH4/H2O/O2 provides an indication that a structural situation is achieved that remains stable with cycle repetition.

3. Discussion

It is mentioned in the literature that an intrinsic mixed oxidation state of Mn (Mn4+/Mn3+) or even Mn2+ exists in pure LaMnO3 perovskites through the so-called discommendation reaction [44,45]. These considerations are important when the material attains excess oxygen under oxidative conditions. In the experiments reported in this article, all materials start the reduction step of the looping cycles after being equilibrated at 900 °C in a He atmosphere, and it is not likely that they contain excess oxygen. Instead, they approach stoichiometry [46]. It is, therefore, assumed that LaMnO3 contains only Mn in its reference valence 3+  M n M n × . In the following, the Kroger–Vink notation will be used for all point defects involved in the discussion [47].
When Ca2+ substitutes for La3+ in A sites, the defect C a L a is created. The extra negative charge of this defect is not compensated by oxygen vacancies V O but by the oxidation of one Mn3+ to Mn4+, leading to the compensating defect M n M n [43]. This is in qualitative agreement with the cell shrinkage as a function of the Ca content shown in Figure 2, as well as with the transformation of the rhombohedral structure to orthorhombic upon the introduction of Ca due to increasing Mn4+ content and reduction of Jahn–Teller distortions caused by Mn3+. If we assume that per chemical formula, fraction x of La3+ ions is substituted by Ca2+ ions, then the material formula could be written in the following form:
L a L a , 1 x × C a L a , x M n M n , 1 x × M n M n , x O 3
The concentration of M n M n is extrinsically determined by the dopant through the neutrality condition C a L a = M n M n .
Upon the introduction of CH4, the initial overall reaction seems to be
C H 4 + 2 O 2 C O 2 + 2 H 2 O
It is believed that the oxygen that takes place in this reaction is lattice oxygen that leaves the material, while Mn4+ is reduced back to Mn3+, with the simultaneous creation of oxygen vacancies, which is described by the following equation:
O O × + 2 M n M n V O + 2 M n M n × + 1 2 O 2
Assuming that fraction δ of oxygen vacancies are created per chemical formula, then the material formula in (1) can be modified to the following:
L a L a , 1 x × C a L a , x M n M n , 1 x 2 δ × M n M n , x 2 δ O 3 δ V O , δ
And by a combination of (2) and (3), the final CH4 reaction with the material can be written as follows:
C H 4 + 4 O O × + 8 M n M n 4 V O + 8 M n M n × + C O 2 + 2 H 2 O
Therefore, the higher the Ca content, the higher the Mn4+ content, and, consequently, the higher the oxygen deficiency, causing reaction (2) to proceed, which is in agreement with the results shown in Figure 3. It is believed that the initial linear part of the oxygen deficiency evolution per pulse reflects the exact reaction described by (5). This linear part is extended to higher deficiencies as the Ca content increases. The dependency of CH4 conversion as a function of Ca content, as shown in Figure 4, can also be explained. The higher the Mn4+ concentration, the more oxygen may become available and, consequently, the higher the conversion. Finally, the results shown in Figure 5 indicate that this initial CH4 reaction with lattice oxygen is directed to CO2 and H2O, in agreement with (2).
As the reduction step proceeds and Mn4+ is depleted, then another reaction starts to gradually become of importance, which leads to the production of H2 and CO. In Figure 3, this is indicated by the gradual deviation of the oxygen deficiency evolution from the initial linear part and in Figure 5 by the gradual appearance of CO and H2 in the products with parallel disappearance of CO2 and H2O. The Mn4+ is gradually depleted so that at the end, the material formula given by (4) can be modified to
L a L a , 1 x × C a L a , x M n M n × O 3 x 2 V O , x 2
The charge of the dopant is no longer compensated by Mn4+ but by oxygen vacancies, and the electrical neutrality is given by C a L a = 2 V O .
It is believed that the oxygen required for this second reaction is lattice oxygen delivered by the material though a Mn3+ to Mn2+ reduction, with simultaneous creation of oxygen vacancies as well, which is described by the following equation:
O O × + 2 M n M n x V O + 2 M n M n + 1 2 O 2
Assuming the creation of a fraction n oxygen vacancies per chemical formula though the reduction described by (7), then Formula (6) can be modified to
L a L a , 1 x × C a L a , x M n M n , 1 2 n × M n M n , 2 n O 3 x 2 n V O , x 2 + n
It is, therefore, stated that the Mn3+→Mn2+ consumes lattice oxygen and produces Mn2+ and oxygen vacancies. This statement, however, does not explain the fact that a constant CH4 conversion level is achieved corresponding to constant CO and H2 production per pulse. Consumption of all available oxygen should either lead to total conversion elimination or to structural deterioration; neither of the two is observed. The total conversion elimination indeed occurred when, in separate experiments, the reactants were totally free of oxygen. On the other hand, the stability of the pulse number of all products and reactants, including material oxygen deficiency, has also been observed in experiments with more than 25 CH4 pulses. It is worthwhile to mention that during the course of the entire cycle, no carbon deposition is observed, and all mass balances can be closed with very high accuracy. All previous observations point to the possibility that certain oxygen traces might exist in the He feed bottle, and these might have been responsible for the final constant CH4 conversion.
The total reactions that take place in this phase can be described by the following equations:
C H 4 + 3 2 O 2 C O + 2 H 2 O
C H 4 + 1 2 O 2 C O + 2 H 2
It is, however, not probable that H2 coexists even with traces of oxygen at 900 °C, and it is, therefore, believed that reaction (9) dominates the fuel oxidation at high deficiencies.
Reaction (9) in combination with (7) leads to the following equation:
C H 4 + 3 O O × + 6 M n M n x 3 V O + 6 M n M n + C O + 2 H 2 O
As indicated in Figure 5, no water is observed in the products. When Mn3+→Mn2+ reduction occurs, the material becomes active for water splitting. Consequently, the H2O produced through (9) is dissociated to H2 and O2 through the reverse of Equation (7), which, combined with the water dissociation reaction, leads to the following equation:
H 2 O + V O + 2 M n M n O O × + 2 M n M n x + H 2
As reduction through Equation (11) proceeds, vacancies are created, which are consequently annihilated through Equation (12). At a certain oxygen deficiency, the rate of vacancy creation becomes equal to the rate of vacancy annihilation, and a certain CH4 conversion is achieved without significantly affecting the oxygen deficiency of the material. During the previously mentioned equilibrium, the concentrations of all reaction products were constant. Assuming that all μmoles of H2 are produced from the water dissociation reaction, a H2O conversion can be calculated. The results are shown in Figure 11. H2O conversions as high as 100% could be achieved at oxygen deficiency levels at which the material becomes, through Equation (11), active for water splitting. The deficiency level at which this occurs increases as the Ca content increases.
During the water-splitting step, reaction (11) takes place. The oxygen vacancies are annihilated through the Mn2+→Mn3+ oxidation. Only the n fraction of vacancies may participate in this reaction, and as long as they are depleted, this partial oxidation of the material occurs through water-splitting stops. Further oxidation (i.e., through Mn3+→Mn4+ oxidation, as described by the reverse of Equation (3)) cannot proceed due to thermodynamic reasons. This n fraction of vacancies is intrinsically determined by the thermodynamics of the Mn3+→Mn2+ reduction and does not depend on the Ca content. In Figure 12, the dependence of the oxygen deficiency (δ) achieved at the end of the reduction step is shown as a function of the Ca content. In comparison, the oxygen recovery during the water-splitting step is shown, which is determined as the difference between the initial and final oxygen deficiency during this step (noted as Δ δ H 2 O in the figure). The dependency of the extrinsically determined δ on the Ca content is much stronger than the weak dependency of the intrinsically determined Δ δ H 2 O . The small differences obtained in Δ δ H 2 O as a function of the Ca content can be attributed to the fact that the thermodynamic equilibrium constant of reaction (7) might depend on the Ca content. Such effects are found for La1−xSrxMnO3 that exhibit defect chemistry similar to La1−xCaxMnO3 [40].
All reactions that take place during the reduction and the water-splitting step are endothermic and are, therefore, favored by increasing temperature. This explains the increased oxygen deficiencies during the reduction step and the increased Δ δ H 2 O during the water splitting (Figure 7), as well as the increased CH4 conversions toward an also increased steady-state conversion at the end of the reduction step (Figure 8).
Pulse experiments with CH4 on Ruddlesden–Popper-type perovskites La2−xMxNiO4 (M = Ca, Sr) have been reported in the literature [48]. Although the CH4 conversion per pulse is not characterized by a clearly decreasing trend, the obtained CO, CO2, H2 and H2O product patterns are very similar to those reported in this article. It is stated that surface oxygen is responsible for the total CH4 oxidation to CO2 and H2O, while lattice oxygen favors partial oxidation to CO and H2. This is different from the results presented in this article, where it is stated that total oxidation also consumes lattice oxygen. Besides the very systematic dependency of H2O and CO2 appearance and evolution on the composition (which is not expected for oxygen chemisorbed on the surface), intermediate material examinations by XRD after 5 or 10 CH4 pulses already indicated a unit cell expansion that implies the removal of lattice oxygen and the creation of oxygen vacancies. Stated in a different way, the increased oxygen deficiency and the associated changes in the valence of manganese lead to an increase in the binding energy of oxygen and, finally, to a change in selectivity. Quite recently, it has been reported in the literature that XRD peak shifts in catalysts in operation that involve oxygen exchange with the environment might be due to the appearance of strains caused by phase segregation phenomena [49]. Although such effects cannot be excluded, the systematic relation of the shift with the extent of reduction or oxidation, the very good reproducibility and the very small amounts of detected secondary phases indicate that such phenomena, if present, are of minor importance in this study. For the LaMnO3 samples, a transformation from rhombohedral to orthorhombic took place.
La0.8N0.2MO3 (N = Al, Ca, M = Co, Ni, Fe, Co) perovskites have also been extensively evaluated for the two-step thermochemical water-splitting reaction [34,50]. Structural stability could be achieved by performing the reduction step at temperatures below 1000 °C. Under isothermal operation at 800 °C, water splitting produced about 5 cm3 (STP) H2/g solid. This value is quite comparable with the 13 cm3 (STP) H2/g solid reported in this article at 1000 °C if one takes into account that the material reduction is forced with CH4.

4. Experimental Section

La1−xCaxMnO3 (x = 0, 0.2, 0.3, 0.4) are chemically prepared by the citrate method, which is described elsewhere [51]. After synthesis, all powder samples are calcined to 1000 °C for 6 h under airflow. X-ray diffraction (XRD) is performed on a Bruker D8-Advance diffractometer with CuKα radiation (λ = 0.154 nm) for the examination of the crystalline structure of the samples during various stages of the CLH process. All powder had comparable morphologies consisting of primary particles in the order of 2–3 μm, the majority of which form larger porous agglomerates up to 20–30 μm [28]. The specific surface area of all specimens was between 4 and 5 m2 g−1.
A U-type microreactor, with outer and inner diameters of 6 and 4 mm, respectively, is used for the reactivity experiments, which have been conducted in pulse mode at a constant temperature of 900 °C and also occasionally at 800 and 1000 °C. This mode has been chosen since it allows a better understanding of the reactions involved from the products obtained per pulse. Approximately 100 mg of material are placed in the reactor, which is brought to the experiment temperature and left to equilibrate under He atmosphere (N50 grade, purity > 99.999%) at a total flow rate of 50 cm3 min−1. Initially, 25 pulses of undiluted CH4 are fed to the reactor through a 100 μL loop valve. The reduction step is subsequently followed by water injections until a point where no variations occur in the oxygen content of the material. The material oxidation is completed by ~100 μL O2 pulses until the material returns to a fully oxidized steady state. The duration of each CH4 or O2 pulse was 2 min plus 1 minute needed for the valve to be filled with the reactant. For comparative reasons, experiments without the water-splitting step are also performed. The reactor outlet stream is directed to a quadrupole mass spectrometer (Pfeiffer Vacuum GmbH, Asslar Germany, Model Omnistar GSD 350), where it is quantitatively analyzed. The mass fractions corresponding to water (m/e: 18, 17, 16), hydrogen (m/e: 2), oxygen (m/e: 32, 16), carbon monoxide (m/e: 28, 12) and carbon dioxide (m/e: 44, 28, 12) are monitored after each pulse during the entire three-step cycle. The mass spectrometer response to each of the reactants and reaction products is calibrated by using pulse injections from either pure substances or calibration mixtures. Prior to the evaluation of the perovskites, it has been experimentally confirmed, either through experiments with an empty reactor or with a reactor that contained inactive α-Al2O3 powder, that no activity exists originating from factors other than those related to the perovskite material. Although under partial pressures of O2, the materials attain excess oxygen and become La1−xCaxMnO3+δ, upon equilibration in He, they become nearly stoichiometric (i.e., δ ≈ 0) [28]. In any case, the equilibrium in He is considered as the reference state L a 1 x C a x M n O 3 ± δ r e f and all stoichiometric parameters δ reported in the following paragraphs are in relation to δref. Τhe stoichiometric parameter δ is calculated from the cumulative amount of atomic oxygen that left the material (as CO or CO2) per mole of solid material. The conversion of CH4 per pulse is defined as the percentage of injected CH4 consumed (in CO and CO2). The exact amount of injected CH4 is determined by the sum of the amounts of unreacted CH4, produced CO and produced CO2. It is also determined by the sum of the amounts of unreacted CH4, produced H2, and produced H2O. No difference larger than 5% is observed, which is considered within the experimental error if one takes into account the uncertainty in integrating the tailoring of the water peak. This result provides a strong indication that carbon deposition, if any, is negligible. In addition, the reaction products during the water-splitting step that followed the reduction step did not contain any CO or CO2 traces detectable by the mass spectrometer. Carbon deposition is, therefore, not considered in the analysis of the results reported in this article.

5. Conclusions

The relation between nonstoichiometric of (La1−xCax)MnO3 (x = 0, 0.2, 0.3, 0.4) perovskites and their catalytic performance in a three-step LCH chemical looping process has been studied. The results were interpreted using point defect-based solid-state chemistry.
The addition of Ca in A-sites creates Mn4+, the subsequent reduction of which is associated with the creation of oxygen vacancies and with the oxidation of CH4 to CO2 and H2O.
Further reduction creates Mn2+, as well as the simultaneous creation of oxygen vacancies and the oxidation of CH4 to CO and H2O. Upon this second reduction, the materials become active for water splitting so that all formed H2O dissociates toward H2 production and oxygen vacancy annihilation.
During the water-splitting step, the material performance does not strongly depend on the Ca content. It is stated that this process is determined intrinsically.
The H2 production increases with increasing temperature. At 1000 °C, the water-splitting step gives rise to the production of ~13 cm3 (STP) H2/g solid.
Under the experimental conditions reported in this article, all materials retain the perovskite structure during all steps of the process and exhibit a constant catalytic behavior after 10 cycles of the three-step CLH process.

Author Contributions

M.M.: experimental work and investigation; A.E.: experimental work and investigation; V.Z.: writing—original draft preparation and supervision. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Data Availability Statement

All data generated or analyzed during this study are included in this published article.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Nejat, P.; Jamehzadeh, F.; Taheri, M.M.; Gohari, M.; Majid, M.Z.A. A global review of energy consumption and policy in the residential sector (with an overview of the top ten CO2 emitting countries). Renew. Sustain. Energy Rev. 2015, 43, 843–862. [Google Scholar] [CrossRef]
  2. Cook, J.; Green, S.; Richardson, M. Quantifying the consensus on anthropogenic global warming in the scientific literature. Environ. Res. Lett. 2013, 8, 24024–24027. [Google Scholar] [CrossRef]
  3. Abuelgasim, S.; Wang, W.; Abdalazeez, A. A brief review for chemical looping combustion as a promising CO2 capture technology: Fundamentals and progress. Sci. Total Env. 2020, 764, 142892. [Google Scholar] [CrossRef]
  4. Lyngfelt, A. Chemical looping combustion: Status and development challenges. Energy Fuel 2020, 34, 9077–9093. [Google Scholar] [CrossRef]
  5. 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]
  6. Das, S.; Biswas, A.; Tiwary, C.S.; Paliwal, M. Hydrogen production using chemical looping technology: A review with emphasis on H2 yield of various oxygen carriers. Int. J. Hydrogen Energy 2022, 47, 28322–28352. [Google Scholar] [CrossRef]
  7. Protassova, L.; Snijkers, F. Recent developments in oxygen carrier materials for hydrogen production via chemical looping processes. Fuel 2016, 181, 215–282. [Google Scholar] [CrossRef]
  8. Adanez, J.; Abad, A.; Garcia-Labiano, F.; Gavan, P.; de Diego, L.F. Progress in chemical-looping combustion and reforming technologies. Prog. Energy Combust. Sci. 2012, 38, 215–282. [Google Scholar] [CrossRef]
  9. Mattisson, T.; García-Labiano, F.T.; Kronberger, B.; Lyngfelt, A.; Adánez, J.; Hofbauer, H. Chemical-looping combustion using syngas as fuel. Int. J. Greenh. Gas. Control 2007, 1, 158–169. [Google Scholar] [CrossRef]
  10. Cormos, C.C. Evaluation of syngas-based chemical looping applications for hydrogen and power co-generation with CCS. Int. J. Hydrogen Energy 2012, 37, 13371–13386. [Google Scholar] [CrossRef]
  11. Gnanapragasam, N.V.; Reddy, B.V.; Rosen, M.A. Hydrogen production from coal using coal direct chemical looping and syngas chemical looping combustion systems. Int. J. Hydrogen Energy 2009, 34, 26066-15. [Google Scholar] [CrossRef]
  12. Cho, W.C.; Lee, D.Y.; Seo, M.W.; Kim, S.D.; Kang, K.S.; Bae, K.K.; Kim, C.H.; Jeong, S.U.; Park, C.S. Continuous operation characteristics of chemical looping hydrogen production system. Appl. Energy 2014, 113, 1667–1674. [Google Scholar] [CrossRef]
  13. Chiesa, P.; Lozz, G.; Malandrino, A.; Romano, M.; Piccolo, V. Three-reactors chemical looping process for hydrogen production. Int. J. Hydrogen Energy 2008, 33, 2233–2245. [Google Scholar] [CrossRef]
  14. 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]
  15. 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 process–A review. Catalysts 2020, 10, 926. [Google Scholar] [CrossRef]
  16. Voitic, G.; Hacker, V. Recent advancements in chemical looping water splitting for the production of hydrogen. RSC Adv. 2016, 6, 98267. [Google Scholar] [CrossRef]
  17. 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]
  18. 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]
  19. Ismail, M.; Liu, W.; Dunstan, M.T.; Scott, S.A. Development and performance of iron based oxygen carriers containing calcium ferrites for chemical looping combustion and production of hydrogen. Int. J. Hydrogen Energy 2016, 41, 4073–4084. [Google Scholar] [CrossRef]
  20. Yüzbasi, N.S.; Kierkowska, A.; Müller, C. Development of Fe2O3-based, Al2O3-stabilized oxygen carriers using sol-gel technique for H2 production via chemical looping. Energy Procedia 2017, 114, 436–445. [Google Scholar] [CrossRef]
  21. O′Connell, M.; Norman, A.K.; Hüttermann, C.F.; Morris, M.A. Catalytic oxidation over lanthanum-transition metal perovskite materials. Catal. Today 1999, 47, 123–132. [Google Scholar] [CrossRef]
  22. Mihai, O.; Chen, D.; Holmen, A. Catalytic consequence of oxygen of lanthanum ferrite perovskite in chemical looping reforming of methane. Ind. Eng. Chem. Res. 2011, 50, 2613–2621. [Google Scholar] [CrossRef]
  23. Wei, H.J.; CaO, Y.; Ji, W.J.; Au, C.T. Lattice oxygen of La1−xSrxMO3 (M = Mn, Ni) and LaMnO3−αFβ perovskite oxides for the partial oxidation of methane to synthesis gas. Catal. Commun. 2008, 9, 2509–2514. [Google Scholar] [CrossRef]
  24. Nalbandian, L.; Evdou, A.; Matsouka, C.; Zaspalis, V. Assessment of (La1−xSrx)MnO3±δ perovskites as oxygen carrier materials in chemical-looping processes. Fuel Process. Technol. 2022, 226, 107086. [Google Scholar] [CrossRef]
  25. He, F.; Li, X.; Zhao, K.; Huang, Z.; Wei, G.; Li, H. The use of La1−xSrxFeO3 perovskite type oxides as oxygen carriers in chemical-looping reforming of methane. Fuel 2013, 108, 465–473. [Google Scholar] [CrossRef]
  26. Zhao, K.; Huang, Z.; Li, H.; Zhao, Z. LaFe1−xNixO3 perovskites as oxygen carriers for chemical-looping steam methane reforming. J. Eng. Thermophys. 2014, 35, 2546–2550. [Google Scholar]
  27. Zhao, K.; He, F.; Huang, Z.; Wei, G.; Zheng, A.; Li, H.; Zhao, Z. Perovskite-type oxides LaFe1−xCoxO3 for chemical looping steam methane reforming to syngas and hydrogen co-production. Appl. Energy 2016, 168, 168–193. [Google Scholar] [CrossRef]
  28. Evdou, A.; Georgitsis, T.; Matsouka, C.; Pachatouridou, E.; Iliopoulou, E.; Zaspalis, V. Defect Chemistry and Chemical Looping Performance of La1−xMxMnO3 (M = Sr, Ca, (x = 0–0.5)) Perovskites. Nanomaterials 2022, 12, 3461. [Google Scholar] [CrossRef] [PubMed]
  29. Liang, H. Study on the effect of CeO2 on Fe2O3/LaNiO3 as the oxygen carrier applied in chemical-looping hydrogen generation. Int. J. Hydrogen Energy 2015, 40, 13338–13343. [Google Scholar] [CrossRef]
  30. 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]
  31. 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. Sustain. Chem. Eng. 2013, 1, 364–373. [Google Scholar] [CrossRef]
  32. Zhao, K.; He, F.; Huang, Z.; Zheng, A.; Li, H.; Zhao, Z. Three-dimensionally ordered macroporous LaFeO3 perovskites for chemical-looping steam reforming of methane. Int. J. Hydrogen Energy 2014, 39, 3243–3252. [Google Scholar] [CrossRef]
  33. Pérez, A.; Orfila, M.; Linares, M.; Sanz, R.; Marug, J.; Molina, R.; Botas, J.A. Hydrogen production by isothermal thermochemical cycles using La0.8Ca0.2MeO3±δ (Me = Co, Ni, Fe and Cu) perovskites. Int. J. Hydrogen Energy 2024, 52, 1101–1112. [Google Scholar] [CrossRef]
  34. Orfila, M.; Linares, M.; Pérez, A.; Barras-García, I.; Molina, R.; Marugán, J.; Botas, J.A.; Sanz, R. Experimental evaluation and energy analysis of a two-step water splitting thermochemical cycle for solar hydrogen production based on La0.8Sr0.2CoO3−δ perovskite. Int. J. Hydrogen Energy 2022, 47, 41209–41222. [Google Scholar] [CrossRef]
  35. Murray, E.P.; Barnett, S.A. (LaSr)MnO3-(Ce,Gd)O2−x composite cathodes for solid oxide fuel cells. Solid. State Ion. 2002, 143, 265–273. [Google Scholar] [CrossRef]
  36. Haile, S.M. Fuel cell materials and components. Acta Mater. 2003, 52, 5981–6000. [Google Scholar] [CrossRef]
  37. Shen, M.; Zhao, Z.; Chen, J.; Su, Y.; Wang, J.; Wang, X. Effects of calcium substitute in LaMnO3 perovskites for NO catalytic oxidation. J. Rare Earths 2013, 31, 119–123. [Google Scholar] [CrossRef]
  38. Tulloch, J.; Donne, S.W. Activity of perovskite La1−xSrxMnO3 catalysts towards oxygen reduction in alkaline electrolytes. J. Power Sources 2009, 188, 359–366. [Google Scholar] [CrossRef]
  39. Hallberg, P.; Källén, M.; Jing, D.; Snijkers, F.; van Noven, J.; Rvdén, M.; Lyngfelt, A. Experimental investigation of CaMnO3−δ based oxygen carriers used in continuous chemical-looping combustion. Int. J. Chem. Eng. 2014, 2014, 412517. [Google Scholar] [CrossRef]
  40. Nowotny, J.; Rekas, M. Defect chemistry of (La,Sr)MnO3. J. Amer. Ceram. Soc. 1998, 81, 67–80. [Google Scholar] [CrossRef]
  41. Zou, D.; Yi, Y.; Song, Y.; Guan, D.; Xu, M.; Ran, R.; Wang, W.; Zhou, W.; Shao, Z. BaCe0.16Y0.04Fe0.8O3−δ nanocomposite: A new high-performance cobalt-free triple conducting cathode for protonic ceramic fuel cells operating at reduced temperatures. J. Mater. Chem. A 2022, 10, 5381–5390. [Google Scholar] [CrossRef]
  42. Gellings, P.J.; Bouwmeester, H.J.M. Ion and mixed conducting oxides as catalysts. Catal. Today 1992, 12, 1–101. [Google Scholar] [CrossRef]
  43. Heuer, S.A.; Schierholz, R.; Alekseev, E.V.; Peters, L.; Mueller, D.N.; Duchoň, T.; Vibhu, V.; Tempel, H.; De Haart, L.G.J.; Eichel, R.A. Oxygen nonstoichiometry and valance state of manganese in La1−xCaxMnO3+δ. ACS Omega 2021, 6, 9638–9652. [Google Scholar] [CrossRef] [PubMed]
  44. Töpfer, J.; Goodenough, J.B. LaMnO3+δ revisited. J. Sol. St. Chem. 1997, 130, 117–128. [Google Scholar] [CrossRef]
  45. Zuev, A.Y.; Tsvetkov, D.S. Oxygen nonstoichiometry, defect structure and defect-induced expansion of undoped perovskite LaMnO3±δ. Solid. State Ion. 2010, 181, 557–563. [Google Scholar] [CrossRef]
  46. Wold, A.; Arnott, R.J. Preparation and crystallographic properties of the systems LaMn1−xMnxO3+λ and LaMn1−xNixO3+λ. J. Phys. Chem. Solids 1959, 9, 176–180. [Google Scholar] [CrossRef]
  47. Kröger, F.A.; Vink, H.J. Solid State Physics 3; Setiz, F., Turnbull, D., Eds.; Academic Press: New York, NY, USA, 1956; p. 307. [Google Scholar]
  48. Melo, V.R.M.; Medeiros, R.L.B.A.; Braga, R.M.; Macedo, H.P.; Ruiz, J.A.C.; Moure, G.T.; Melo, M.A.F.; Melo, D.M.A. Study of the reactivity of double-perovskite type oxide La2−xMxNiO4 (M=Ca or Sr) for chemical looping hydrogen production. Int. J. Hydrogen Energy 2018, 43, 1406–1414. [Google Scholar] [CrossRef]
  49. Guan, D.; Shi, C.; Xu, H.; Gu, Y.; Zhong, J.; Sha, Y.; Hu, Z.; Ni, M.; Shao, Z. Simultaneously mastering operando strain and reconstruction effects via phase-segregation strategy for enhanced oxygen-evolving electrocatalysis. J. Energy Chem. 2023, 82, 572–580. [Google Scholar] [CrossRef]
  50. Pérez, A.; Orfila, M.; Linares, M.; Sanz, R.; Marugán, J.; Molina, R.; Botas, J.A. Hydrogen production by thermochemical water splitting with La0.8Al0.2MeO3−δ (Me = Fe, Co, Ni and Cu) perovskites prepared under controlled pH. Catal. Today 2022, 390, 22–33. [Google Scholar] [CrossRef]
  51. Nalbandian, L.; Evdou, A.; Zaspalis, V. La1−xSrxMyFe1−yO3−δ perovskites as oxygen-carrier materials for chemical looping reforming. Int. J. Hydrogen Energy 2011, 36, 6657–6670. [Google Scholar] [CrossRef]
Figure 1. Simplified schematic flow diagrams of the (a) CLC/CLR and (b) CLH processes.
Figure 1. Simplified schematic flow diagrams of the (a) CLC/CLR and (b) CLH processes.
Catalysts 14 00431 g001
Figure 2. X-ray diffraction patterns of the “as prepared” La1−xCaxMnO3 samples.
Figure 2. X-ray diffraction patterns of the “as prepared” La1−xCaxMnO3 samples.
Catalysts 14 00431 g002
Figure 3. The evolution of the stoichiometric parameter δ during the course of the three-step CLH cycle. Insert: Idem for a CL cycle without the water-splitting step.
Figure 3. The evolution of the stoichiometric parameter δ during the course of the three-step CLH cycle. Insert: Idem for a CL cycle without the water-splitting step.
Catalysts 14 00431 g003
Figure 4. CH4 conversion per pulse (open symbols, left axis) and cumulative methane consumption (filled symbols, right axis) during the reduction step as a function of the pulse number.
Figure 4. CH4 conversion per pulse (open symbols, left axis) and cumulative methane consumption (filled symbols, right axis) during the reduction step as a function of the pulse number.
Catalysts 14 00431 g004
Figure 5. Cumulative production of (a) CO2 (filled symbols, left axis) and CO (open symbols, right axis) and (b) H2O (filled symbols, left axis) and H2 (open symbols, right axis) as a function of the pulse number for different Ca contents.
Figure 5. Cumulative production of (a) CO2 (filled symbols, left axis) and CO (open symbols, right axis) and (b) H2O (filled symbols, left axis) and H2 (open symbols, right axis) as a function of the pulse number for different Ca contents.
Catalysts 14 00431 g005
Figure 6. Cumulative H2 production during the water-splitting step as a function of the pulse number for different Ca contents.
Figure 6. Cumulative H2 production during the water-splitting step as a function of the pulse number for different Ca contents.
Catalysts 14 00431 g006
Figure 7. The evolution of the stoichiometric parameter δ during the reduction and the water-splitting step as a function of the pulse number for (La0.7Ca0.3)MnO3 and (La0.6Ca0.4)MnO3 perovskites at three different operation temperatures.
Figure 7. The evolution of the stoichiometric parameter δ during the reduction and the water-splitting step as a function of the pulse number for (La0.7Ca0.3)MnO3 and (La0.6Ca0.4)MnO3 perovskites at three different operation temperatures.
Catalysts 14 00431 g007
Figure 8. CH4 conversion (filled symbols, left axis) and cumulative H2 production (open symbols, right axis) during the reduction and the water-splitting step as a function of the pulse number for (La0.7Ca0.3)MnO3 and (La0.6Ca0.4)MnO3 perovskites at three different operation temperatures.
Figure 8. CH4 conversion (filled symbols, left axis) and cumulative H2 production (open symbols, right axis) during the reduction and the water-splitting step as a function of the pulse number for (La0.7Ca0.3)MnO3 and (La0.6Ca0.4)MnO3 perovskites at three different operation temperatures.
Catalysts 14 00431 g008
Figure 9. Oxygen deficiency (δ) after each process step during 10 successive reduction–partial oxidation–total oxidation cycles as a function of the cycle number.
Figure 9. Oxygen deficiency (δ) after each process step during 10 successive reduction–partial oxidation–total oxidation cycles as a function of the cycle number.
Catalysts 14 00431 g009
Figure 10. XRD spectra of (a) (La0.7Ca0.3)MnO3−δ as prepared and after each step of the three steps of the CLH process and (b) (La0.6Ca0.4)MnO3−δ perovskites after each step of the CLH process and after CLR or CLH cycle repetition.
Figure 10. XRD spectra of (a) (La0.7Ca0.3)MnO3−δ as prepared and after each step of the three steps of the CLH process and (b) (La0.6Ca0.4)MnO3−δ perovskites after each step of the CLH process and after CLR or CLH cycle repetition.
Catalysts 14 00431 g010
Figure 11. Water conversion during the reduction step as a function of the oxygen deficiency for materials with different Ca contents.
Figure 11. Water conversion during the reduction step as a function of the oxygen deficiency for materials with different Ca contents.
Catalysts 14 00431 g011
Figure 12. Oxygen deficiencies (δ) during the reduction step and deficiency difference Δ δ H 2 O during the water-splitting step as a function of the Ca content.
Figure 12. Oxygen deficiencies (δ) during the reduction step and deficiency difference Δ δ H 2 O during the water-splitting step as a function of the Ca content.
Catalysts 14 00431 g012
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

Moschos, M.; Evdou, A.; Zaspalis, V. (La1−xCax)MnO3−δ (x = 0, 0.2, 0.3, 0.4) Perovskites as Redox Catalysts in Chemical Looping Hydrogen Production Process: The Relation between Defect Chemistry and Redox Performance. Catalysts 2024, 14, 431. https://doi.org/10.3390/catal14070431

AMA Style

Moschos M, Evdou A, Zaspalis V. (La1−xCax)MnO3−δ (x = 0, 0.2, 0.3, 0.4) Perovskites as Redox Catalysts in Chemical Looping Hydrogen Production Process: The Relation between Defect Chemistry and Redox Performance. Catalysts. 2024; 14(7):431. https://doi.org/10.3390/catal14070431

Chicago/Turabian Style

Moschos, Moschos, Antigoni Evdou, and Vassilios Zaspalis. 2024. "(La1−xCax)MnO3−δ (x = 0, 0.2, 0.3, 0.4) Perovskites as Redox Catalysts in Chemical Looping Hydrogen Production Process: The Relation between Defect Chemistry and Redox Performance" Catalysts 14, no. 7: 431. https://doi.org/10.3390/catal14070431

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