Preprint
Review

Withdrawn:

Alternative Oxide-Based Materials for Electrochemical and Catalytic Applications: A Review

Altmetrics

Downloads

211

Views

115

Comments

0

This preprint has been withdrawn

Submitted:

16 August 2024

Posted:

20 August 2024

Withdrawn:

09 December 2024

Alerts
Abstract
Recent trends concerning hydrogen as a fuel of the future and problems related to its production, storage and utilization increase the demand for efficient technologies related to these problems. A large part of such technologies is related to application of advanced materials based on various complex oxides which are used as efficient catalysts for fuel transformation into hydrogen and syngas, materials for membranes with selective permeability for hydrogen and syngas production, components of solid oxide fuel cells/electrolyzers as well as many other applications in energy, electrochemistry and catalysis. Such materials are typically based on oxides and solid solutions with perovskite and fluorite structures or their derivatives. Attention is also paid to the materials based on spinel, apatite, mayenite, sheelite, fergusonite structures. The above-mentioned materials are continuously studied in order to modify and optimize their functional properties. On the other hand, other classes of materials (weberites, monazites, etc.) are recently discovered and considered as candidate materials for the aforementioned applications due to their specific functional properties or a great potential of their improvement. In this review, some of these materials are considered.
Keywords: 
Subject: Chemistry and Materials Science  -   Ceramics and Composites

1. Introduction

Ceramic and cermet materials based on various oxides, solid solutions and composites have found wide applications in energy, electrochemistry and catalysis, such as materials for components of solid oxide fuel cells (SOFC) and electrolyzers (SOE) (electrodes, electrolytes, interconnects) [1,2,3,4,5,6,7,8,9,10,11,12], oxygen and hydrogen separation membranes [10,11,12,13,14,15,16,17,18,19,20], catalysts for fuel transformation into hydrogen and syngas or their supports [11,12,19,20,21,22,23,24,25,26,27,28,29,30,31,32,33,34,35,36,37,38,39], as well as various applications in different fields such as sensors [12,40,41,42], electronics [12,43,44] and many others [11,12,14,15,45,46,47].
Recent reviews devoted to the materials for SOFC/SOE components, permselective membranes, various catalysts and their supports are generally focused on the ionic (oxide ionic and/or protonic) and mixed ionic-electronic (including H+/O2-/e- triple conductive) conducting oxides and solid solutions with perovskite structure [1,2,3,4,5,6,7,8,9,19,24,25,27,33,36,39,48,49,50,51,52,53,54,55,56,57,58,59,60,61,62,63] and its derivatives such as Ruddlesden – Popper phases [5,6,7,8,9,15,27,36,48,49,51,53,56,58,59,61,62,63,64,65,66], double perovskites [1,2,5,6,7,8,9,19,27,36,48,49,50,51,53,56,57,58,59,61,62,64,67,68], brownmillerites [8,61,64,69], with fluorite structure [1,2,3,6,7,8,9,15,19,20,24,25,32,33,34,37,38,49,52,53,61,62,63,64,70,71] and its derivatives such as bixbyites [8,52,61,62], pyrochlores [2,8,20,36,39,61,62,63,64,72], weberites [61,72], and various (nano)composites (including core-shell nanocomposites) based on them [1,5,6,7,8,9,13,15,19,20,24,25,27,32,33,34,36,37,39,48,49,50,51,56,61,62,63]. Oxides with tetragonal moieties such as sheelites [8,61,64,73,74], fergusonites [64,74], monazites [64,75], melitites [8,73], LAMOX [3,8,52,61,73], apatites [3,8,61,64], mayenites [3,8,37,39,61,64,73] have attracted attention as promising oxygen and proton conducting materials due to the deformation and rotation flexibility of tetrahedral units and other transport features such as various types of cooperative transport mechanisms. Spinels and composites based on them were widely applied as well [3,5,15,19,20,24,25,32,33,34,36,56,61,62,63]. Figure 1 summarizes this information.
There have been other types of materials, often recently discovered ones, which are not frequently mentioned in reviews, however, they have displayed a good specific functional properties or have possessed a great prospective of their improvement after variations of their composition. These materials have included many types of non-cubic structures for rare earth element (REE) tungstates and molybdates [50,52,76,77,78], various layered oxides such as Ca, Ba and Li cobaltites [2,3,6,37,79,80] and others [3,37,39,78,81]. In this work, selected types of such uncommon materials have been mainly reviewed.

2. New Non-Cubic Structural Types of Oxygen Ion and Proton Conductors

It has been conventionally assumed that cubic structures, specifically perovskites, fluorites, and pyrochlores, represent the most promising materials for oxygen separation membranes. Cubic structures are typically formed at high temperatures, whereas modifications with a lower symmetry (e.g., rhombic, hexagonal, tetragonal, and monoclinic) are obtained at lower temperatures. Due to this, in order to obtain cubic modifications that are stable at low temperatures, e.g., ZrO2, doping with Y, Sc, Ln cations has been carried out. Y-stabilized zirconia (9 mol. % Y2O3) (YSZ) with the fluorite structure has been the most known and practically applied solid electrolyte in modern SOFC. However, its functionalily at high temperatures ~800 – 1000 °C as an electrolyte as well as a constituent of Ni/ZrO2 SOFC anode has been constrained by a discernible decline in oxide ionic conductivity. The oxide ionic conductivity exhibits a decrease of ~40 % at 950 °C after 2500 h of operation [82]. A contributing factor to the degradation is the partial phase transition into the low-temperature tetragonal modification.
However, accumulation of the experimental results revealed that other materials with less symmetric structures can demonstrate high values of oxide ionic and protonic conductivity. In light of these findings, non-cubic polymorphic modifications with intrinsic ionic conductivity stable within the temperature range of 500–800 °C have attracted significant interest.

2.1. REE Molybdates and Tungstates

Systems La2O3–MO3 (M = Mo, W) have been known by the fact that highly-conducting oxide ionic conductor La2Mo2O9 (LAMOX) [83] has been firstly discovered, and then solid solutions La6–xWO12–δ (x = 0.3–0.7) with sufficiently high conductivity to be used as proton-conducting SOFC electrolytes [84,85,86] and proton-conducting membranes for hydrogen separation [87,88,89] have been revealed.
It is noteworthy that the specified compounds possess a rather unusual structure characterized by the large size unit cells with a complex arrangement. La2Mo2O9 demonstrates first-order phase transitions at 580 °C [90]: low-temperature monoclinic α-phase (P21) undergoes a transformation into the high-temperature cubic β-phase (P213) with increasing the conductivity by two orders of magnitude (0.06 S cm–1 at 800 °C). It is known that, e.g., the low-temperature monoclinic α-phase of LAMOX contains 312 atoms within the single unit cell. However, at room temperature after cooling, metastable modification of the β-phase, designated as the β՜-phase, partially preserves along with the α-modification which is attributed to the difficulty in arranging too large number of oxygen atoms within the unit cell of the low-temperature α-phase [91]. It has been demonstrated that the splitting of one of the three main positions of oxygen is present in the β՜-structure, which characterizes it as partially ordered [92]. Therefore, a mixture of α+β՜-phases is present at room temperature. It is evident that the connection between the highly conductive cubic phases and their low-temperature polymorphs significantly impacts the practical applications of highly conductive cubic electrolytes.
A search for highly conductive materials amongst non-cubic molybdates Ln2MoO6 (Ln = La, Pr, Nd, Sm, Gd, Dy) has been carried out during last years [93,94]. The oxygen pressure-dependent conductivity isotherms of Ln2MoO6 (Ln = La, Pr, and Nd) at 1 × 10−18 – 0.21 atm pO2 indicate the presence of a mixed ionic and electronic conductivity within the temperature range of 700−900 °C [93]. The oxymolybdates with large cations (La, Pr, and Nd) possess a tetragonal symmetry. The highest ionic component of the conductivity, 10−3 S cm–1 at 800 °C, has been discovered for Pr2MoO6 possessing a space group I4̅c2. All three layered light REE oxymolybdates Ln2MoO6 (Ln = La, Pr, and Nd) exhibit the proton conductivity in a wet atmosphere.
The conductivity of the monoclinic polymorphs of Ln2MoO6 (Ln = Sm, Gd, Dy) oxymolybdates has been investigated by theoretical and experimental methods [94]. All single-phase compounds are isostructural with a monoclinic symmetry. It is noteworthy that the ionic component of conductivity demonstrated an increase with decreasing the ionic radius in the series of molybdates: ~10–5 S cm–1 for Sm2MoO6, ~10–4 S cm–1 for Gd2MoO6 and ~10–3 S cm–1 for Dy2MoO6 at 800 °C according to impedance spectroscopy data. The oxygen pressure isotherms in Ln2MoO6 (Ln = Sm, Gd, Dy) samples indicated the electronic conductivity contribution in the temperature range of 500–800 °C at high partial oxygen pressures. Hence, the increase in the ionic component of the conductivity with decreasing lanthanide ionic radius in the series of the monoclinic oxymolybdates Ln2MoO6 (Ln = Sm, Gd, Dy) has been observed (Figure 2a) [94]. The oxygen migration barrier for Ln2MoO6, as calculated by means of the Density-Functional Theory (DFT) using Nudged Elastic Band (NEB) method, exhibits a decrease in the Ln row with a decrease in the ionicity degree of the Ln–O bond, as follows: La > Pr > Nd > Sm > Gd > Dy. The lowest energy barriers of oxygen migration in Ln2MoO6 (Ln = Sm, Gd, Dy) were observed for the O1–O1, O1–O2, O6–O6 and O4–O5 transitions. The DFT – NEB oxygen diffusion maps acquired according to these calculations are similar for Ln2MoO6 (Ln = Sm, Gd, Dy). For illustrative purposes, a map for Ln = Dy is provided in Figure 2b [94].
The objective of the work co-authored by Qin Li, and Venkataraman Thangadurai [95] was to develop new mixed conductors that would possess the chemical-physical properties of Ni–YSZ while avoiding the issues inherent to the conventional cermet anodes. Modern anode disadvantages that lead to a decline in SOFC operation include a decrease in porosity, the destruction of the cermet microstructure due to carbon deposition, and the formation of nickel sulfide during sulfur adsorption [96]. Furthermore, metal-based anodes tend to suffer from the particle agglomerations during the operation, which results in shortening the effective triple-phase boundary (TPB). TPB is an electrochemically active region where the contact of a solid electrolyte, a fuel (in the case of the anode) or oxygen (in the case of the cathode) molecule and an electrode occurs. In the aforementioned work, the authors have explored monoclinic structure of Sm2‒xAxM1–yByO6−δ (A = Ca, Sr; M = Mo, W; B = Ce, Ni) solid solutions as precursors for the development of solid oxide fuel cells anodes. Among the samples studied, Sm1.8Ca0.2MoO6−δ (SMO) displays a total conductivity of 0.12 S cm–1 at 550 °C in wet H2 with an activation energy of 0.06 eV. Ca-doped SMO appears to be chemically stable against reaction with YSZ electrolyte at 800 °C for 24 h in wet H2.
For the solid solutions based on lanthanum tungstate La6–xWO12–δ (x = 0.2–1) Magrasó et al. [86] proposed the formula La28–xW4+xO54+1.5x( V O )2–1.5x (x is W on La positions), with an increased unit cell, which ensured better agreement between the measured and X-ray densities than did the previously used formula La6–xWO12–δ (x = 0.2–1). Note that the tungstates also have a large unit cell with a complex structure, like α-La2Mo2O9. Hence, large and complex unit cells are not rare in the La2O3–MO3 (M = Mo, W) systems. This feature is possibly due to the considerable ion size mismatch between La and Mo(W).
Along with the best La6–xWO12–δ (x = 0.3–0.7) proton conductors, the La2WO6 compound [97,98,99] and other compounds of this system – La18W10O57 [100], La6W2O15 [101], and La10W2O21 [102] – attracted researchers’ attention at the same time (Figure 3). A number of studies were concerned with the structure and ionic conductivity of these compounds [103,104,105,106,107]. γ-La6W2O15 [106] and β-La2WO6 [104,105] were shown to have proton conductivity, however, it was two orders of magnitude lower than that of La6–xWO12–δ (x = 0.3–0.7).
Further studies of the compounds in the system Ln2O3–WO3 (Ln = Nd – Yb) enabled the authors of the work [108] to identify a novel class of oxide ionic conductors Ln14W4O33 with pseudorhombohedral structure. Reliable description of this structure has still not been accomplished up to now.
Tungstates Ln14W4O33 were divided by McCarthy [109] into three groups. The structure of heavy REE tungstates Ln14W4O33 (where Ln = Er–Lu) is the most understandable, corresponding to a simple rhombohedral cell. The structure of Ln14W4O33 (Ln = Ho, Y) is particularly complicated due to the splitting of the strong reflections (214), (431), (422) of a simple rhombohedral cell and has not been completely solved. Ln14W4O33 (Ln = Nd–Dy) is even more problematic, as its structure is proposed to be pseudorhombohedral, with only 4 non-splitted reflections. The study of the structure continued later, and in 1996 the authors [110] proposed a body-centered monoclinic unit cell for Gd14W4O33, but further study of this compound by electron diffraction and refinement of the data by the X-ray diffraction method (with Rietveld refinement) led the authors to conclude that the Gd14W4O33 unit cell has a larger size and is not fully described within the framework of the proposed model. Thus, for the tungstates Ln14W4O33 (Ln = Nd–Dy) and Ln14W4O33 (Ln = Ho, Y) the structure has not been completely resolved even today. Probably, in this case a specific structure forms as well which possesses similar to lanthanum tungstates and molybdates giant cell up to ∼8800 Å3 in size [86,111,112,113], which is structurally closer to short-chain proteins than to ordinary inorganic materials [111].
Ln14W4O33 (Ln = Nd, Sm, Gd) tungstates have a wide electrolytic range (10–15–1 atm, Figure 4) above 700 °C. It should be noted that under oxidizing conditions the contribution of hole conductivity is absent in these compounds. Among the pseudorhombohedral phases of Ln14W4O33 (Ln = Nd, Sm, Gd), Nd14W4O33 exhibits the highest oxygen-ion conductivity (4 × 10–4 S cm–1 at 700 °C). The electrolytic domain of the second half of the REE Ln14W4O33 (Ln = Dy, Ho, Er, Tm, Yb) series becomes narrower (10–10–10–5 atm) and at the same time has a slight slope indicating the presence of impurity electronic charge carriers. Investigations in a reducing atmosphere confirmed that the reduction tendency increases with decreasing lanthanide ionic radius in the Ln14W4O33 tungsten series. The variations of the total conductivity in the series of pseudorhombohedral phases Ln14W4O33 (Ln = Dy, Ho, Er, Tm, Yb) are associated with local changes in the short-range order structure, as evidenced by Raman spectroscopy (Figure 5).
In the work [114], a search for the synthesis conditions for various polymorph modifications of the Nd2WO6 ceramics was carried out including application of the technique of long-term isothermal calcinations. The phase formation of neodymium tungstate Nd2WO6 from mechanically activated oxides in a wide temperature range of 25–1600 °С was studied [115]. The conditions of forming various polymorph modifications were determined: low-temperature rhombic (β-Nd2WO6 and δ- Nd2WO6 (P212121 (No. 19)) and high-temperature monoclinic Nd2WO6 (sp. gr. C12/c1 (15)) ones. Although β-Nd2WO6 and δ-Nd2WO6 stabilized in ceramics for a first time belong to the same space group (P212121 (No. 19), their structures and, thus, XRD patterns differ. It was successful in stabilizing δ-Nd2WO6 in a pure form by using long-term isothermal calcination during 100 h at 900 °C. Moreover, due to mechanical activation, the same polymorph was successfully obtained by short-term calcination at 1600 °C for 1 h. For both modifications of neodymium tungstate, rhombic δ-Nd2WO6 and monoclinic Nd2WO6, the protonic conductivity with the effective activation energy of ~1.05 eV was discovered. However, for Ca-containing solid solutions (Nd1–xCax)2WO6-δ (x = 0.01), which total conductivity increases compared to the undoped monoclinic Nd2WO6, hole conductivity predominates in air. Similar to the tetragonal Nd2MoO6 [93], the total conductivity of monoclinic modification of Nd2WO6 is as low as ~ 1 × 10–6 S cm–1 at 550 °С in dry air.
Figure 5. Raman spectra of a series of tungstates Ln14W4O33 (Ln = Nd, Sm, Gd, Dy, Ho, Er, Tm, Yb) [108]. Reprinted from Ceramics International, Vol. 50, Anna Shlyakhtina et al, Impact of Ln cation on the oxygen ion conductivity of Ln14W4O33 (Ln = Nd, Sm, Gd, Dy, Ho, Er, Tm, Yb) tungstates, Pages No. 704-713, Copyright (2024), with permission from Elsevier.
Figure 5. Raman spectra of a series of tungstates Ln14W4O33 (Ln = Nd, Sm, Gd, Dy, Ho, Er, Tm, Yb) [108]. Reprinted from Ceramics International, Vol. 50, Anna Shlyakhtina et al, Impact of Ln cation on the oxygen ion conductivity of Ln14W4O33 (Ln = Nd, Sm, Gd, Dy, Ho, Er, Tm, Yb) tungstates, Pages No. 704-713, Copyright (2024), with permission from Elsevier.
Preprints 115443 g005
In ternary systems Ln2O3 – Bi2O3MO3 (Ln = La, Pr, Nd; M = Mo, W), a rich phase polymorphism was observed as well [116,117,118,119]. The formation of various cubic, tetragonal, pseudotetragonal and monoclinic phases, or their mixture, takes place depending on the molar ratio of cations. Such materials include various compounds, including those with monoclinic Bi3.24La2W0.76O10.14-like, tetragonal β- and β’-Bi2O3-like, Bi14WO24-like and Bi14W2O27-like, Aurivillius-like Bi2MO6, cubic δ-Bi2O3-like, and numerous others structures, some of which remain unresolved. Such compounds demonstrated a high total conductivity (up to ~10–4 – 10–1 S cm–1 at 700 °C). The oxide ionic conductivity contribution is believed to be high, and the protonic conductivity component apparently emerges in some cases [116,117,118,119].
Various REE tungstates and molybdates can exhibit a good catalytic performance when used as catalysts themselves or as catalyst supports for metal-based catalysts in various oxidation reactions generally due to a high oxygen mobility (as discussed above), reactivity of Mo/W active sites and ability of Mo/W and sometimes REE cations to change their charge and form low-valent species cooperatively working with other active sites (including metallic ones in the case of tungstate/molybdate supports). These reactions include dry reforming [120] and oxidative coupling of methane [121], partial and total oxidation of propane [122], selective oxidation of propene and isobutene into acrolein and methacrolein, respectively [121,123], and others [121,124].
Thus, for Ln-Bi-Ni-Co molybdates (Ln = La, Ce, Pr, Sm, Tb) in the partial oxidation of propane into propene and carbon dioxide, it was demonstrated that Ln-containing catalysts show better catalytic activity compared to Ln-free ones. The Ce- and Tb-containing catalysts provided highest propane conversion and, on the other hand, highest selectivity with respect to CO2. The La- and Pr-containing catalysts demonstrated the highest selectivity with respect to propene. The Sm-containing catalyst possessed the lowest activity [122].

2.2. A2M3O12 Family

The compounds with formula A2M3O12 or A2(MO4)3 (A = Y3+, Ln3+/4+ and other 2+, 3+ or 4+ cations, M = Mo6+, W6+, V5+, P5+) is a large family of materials, being a part of rich polymorphism of oxides with various stoichiometry and various functional properties [52,125,126,127,128]. The oxides in A2M3O12 family commonly found the application related to their negative or close-to-zero thermal expansion coefficient [50,129,130,131,132,133]. However, due to their high ionic conductivity, such oxides are recently considered for using as SOFC/SOE electrolytes [132,134,135,136,137], additives for composite cathodes [50,130], permselective membranes [135] and catalysts for various processes [138,139,140].
A2M3O12 oxides generally possess monoclinic or orthorhombic structure depending on A and M cation nature. E.g., Y tungstate, Sc tungstate and molybdate crystallize in the orthorhombic structure [133,139]; the lighter rare earth tungstates crystallize in monoclinic structure (C2/c) whereas the heavy rare earth tungstates do it in the orthorhombic structure (Pcna) [133]. Such oxides possess open-framework structure with corner-shared AO6 and WO4 polyhedral units, and every oxide anion is bound to one A cation and one M cation [129,130,131,133]. Representations of the monoclinic and orthorhombic polymorphs are given in Figure 6.
A2M3O12 oxides are predominately ionic conductors. Some of these oxides can possess mixed ionic-electronic conductivity due to the ability of A-site cations to vary their charge (such as Eu3+/4+). Total ionic transference number is generally close to 1, however, for some compounds (e.g., Ln2W3O12, Ln = La, Sm, Eu, Gd) can be 0.8 or lower at high temperatures [132,136,137]. According to the nowadays knowledge, there are two main charge carriers in A2M3O12 oxides [136,137,141,142,143]:
  • O2– (typical transference number ~0.65–1) which migrates via vacancy and interstitial mechanisms;
  • (MO4)2– (typical transference number ~0–0.35) which migrates via so-called “Rock and Roll” mechanism (can be considered as O2– vehicle mechanism as well).
Figure 6. Structure of Dy2(WO4)3 compound: (a) orthorhombic (Pnma) and (b) monoclinic (P21/c) polymorphs [130]. Reprinted from Journal of Alloys and Compounds, Vol. 990, Natalia Kireeva, Aslan Yu. Tsivadze, Oxide ceramics of A2M3O12 family with negative and close-to-zero thermal expansion coefficients: Machine learning-based modeling of functional characteristics, Pages No. 174356, Copyright (2024), with permission from Elsevier.
Figure 6. Structure of Dy2(WO4)3 compound: (a) orthorhombic (Pnma) and (b) monoclinic (P21/c) polymorphs [130]. Reprinted from Journal of Alloys and Compounds, Vol. 990, Natalia Kireeva, Aslan Yu. Tsivadze, Oxide ceramics of A2M3O12 family with negative and close-to-zero thermal expansion coefficients: Machine learning-based modeling of functional characteristics, Pages No. 174356, Copyright (2024), with permission from Elsevier.
Preprints 115443 g006
“Rock and Roll” mechanism of (MO4)2– anion migration proceeds in the following way. MO4 tetrahedron in regular position reorients and rolls into the interstitial site. Then neighboring MO4 tetrahedron occupies the position previously occupied by early mentioned MO4 tetrahedron. (MoO4)2– anions are demonstrated to be more mobile than (WO4)2– ones facilitating both rotations and hops [133,141,142,143].
Such transport features provide a good ionic conductivity of A2M3O12 oxides (up to ~10–3 S cm–1 at 700 °C, Figure 7) and oxygen self-diffusion coefficients (Figure 8) which can be estimated using Nernst – Einstein equation (up to ~10–9 cm2 s–1) [132,135,136,137]. Further increase in ionic conductivity can be achieved by doping A- and M-sites with various cations. E.g., doping Sm tungstate with small amounts of Ca2+ and Zn2+ led to increase in the ionic conductivity by a factor of 1.5–2 [135]. Further developments of the materials belonging to this family in order to increase oxygen mobility without losing in stability would amplify their electrochemical (e.g. IT SOFC/SOE electrolytes) and catalytic (e.g. efficient and stable to coking catalysts for fuel transformation or their supports) applications.
As with the REE tungstates/molybdates outlined in the preceding section, numerous members of the A2M3O12 family can function as catalysts or their supports in various reactions due to their transport properties, and the capacity of the cations to alter their charge as surface active site features including participation of the surface (Mo/W)–O and (Mo/W)– V O species in redox processes. Such reactions include partial oxidation [144,145], dry and oxy-dry reforming [146] and oxidative coupling of methane [121], and other hydrocarbons [147,148,149]. selective oxidation of propene and isobutene into acrolein and methacrolein, respectively [121,123], reactions of alcohols’ oxidation [147,150] and others [121,124,147,148].
In partial oxidation of methane, the Fe2Mo3O12 catalyst displays a decrease in selectivity with respect to methanol as the temperature rises. The selectivities with respect to CO, CO2, and formaldehyde exhibit a range of temperature dependencies at various pressures (3 – 67 bar) [144]. For the A2Mo3O12 (A = Al, Ga, In, Sc, Cr, Fe) catalysts in partial oxidation of methane at 750 °C and ambient pressure, different values of CH4 conversion and the products’ selectivities were demonstrated depending on the A cation nature. Various metal additives have been demonstrated to enhance CH4 conversion (Table 1) [145]. For dry and oxi-dry reforming of methane over the Ce2W3O12 catalyst, better performance was demonstrated for the oxi-dry reforming reaction. Total oxidation into carbon dioxide is preferable at T < 800 °C, at higher temperatures, partial oxidation into syngas CO + H2 occurs. The partial oxidation of methane into carbon monoxide can be attributed to oxygen species easily moving to the catalyst surface, thereby promoting the methane oxidation [146].

2.3. Composites Based on REE Molybdates and Tungstates

Some works devoted to composite materials in tungstate and molybdate systems with REE are of interest [151,152].
For the first time, an extensive study of composites (100–x) La2Mo2O9x La2Mo3O12 in a wide concentration range (x = 5, 10, 15, 20, 30, 100) was carried out [151]. Presented results show that (100–x) La2Mo2O9x La2Mo3O12 composites have a high ionic conductivity in the range of x =10–15% (Figure 9). The α-La2Mo2O9−La2Mo3O12 composite materials are a new class of highly-conductive materials demonstrating elevated oxide ionic conductivity. In this case, it is to be noted that composites contain low-temperature monoclinic α-phase La2Mo2O9 along with monoclinic La2Mo3O12 phase. Increase in conductivity, oxygen surface exchange constant and diffusion coefficient for the composites compared to the individual oxides is observed (Figure 10) which is due to a segregation of various ions on grains’ surface and formation of a new fluorite-like phase La5Mo3O16 at the La2Mo2O9-La2Mo3O12 interphase. 3D-modeling of the composite microstructure was carried out based on the SEM image analysis data in order to estimate the conductivity of the interphase La5Mo3O16 layer between La2Mo2O9 and La2Mo3O12 grains. The electrical conductivity values for the composite materials calculated from the 3D-modeling of the microstructure and the experimentally measured ones correlate and demonstrate composite effect.
Figure 9. Concentration dependencies of the bulk conductivity in air for (100–x) mol. % La2Mo2O9x mol.% La2Mo3O12 oxides, where * open red circles correspond to conductivity data calculated from the Nernst – Einstein equation using oxygen diffusion coefficient [151]. Reprinted with permission from Ref. [151]. Copyright (2022) American Chemical Society.
Figure 9. Concentration dependencies of the bulk conductivity in air for (100–x) mol. % La2Mo2O9x mol.% La2Mo3O12 oxides, where * open red circles correspond to conductivity data calculated from the Nernst – Einstein equation using oxygen diffusion coefficient [151]. Reprinted with permission from Ref. [151]. Copyright (2022) American Chemical Society.
Preprints 115443 g009
Figure 10. Arrhenius plots for oxygen surface exchange constant (a) and tracer diffusion coefficient (b) for the La2Mo2O9 (1), La2Mo3O12 (2) and La2Mo2O9 (85 mol. %) – La2Mo3O12 (3) samples [151].
Figure 10. Arrhenius plots for oxygen surface exchange constant (a) and tracer diffusion coefficient (b) for the La2Mo2O9 (1), La2Mo3O12 (2) and La2Mo2O9 (85 mol. %) – La2Mo3O12 (3) samples [151].
Preprints 115443 g010
The protonic conductivity of composites in the La2O3–WO3 system was studied using as an example the protonic conductor La14W4O33 (5 × 10–5 S cm–1 at 600 °C), which is two-phased material comprised of anion-deficient fluorite-like phase La10W2O21 (18 wt.%) and rhombic phase γ-La6W2O15 (82 wt.%) [152]. A high Nd content in the composite based on the γ-La6W2O15 leads to forming the solid solutions based on pseudorhombohedral phase for La14–xNdxW4O33 for x = 12 and 14. Nd-containing composites based on the γ-La6W2O15 possess the protonic conductivity which gradually decreases with decreasing Nd content, and transform into the pure pseudorhombohedral phases La14–xNdxW4O33 (x = 12, 14) with oxide-ionic conductivity (Figure 11a). Synthesis at 1500 °C of mechanically activated 5La2O3+2WO3 oxide mixture resulted in forming the mirror composite including 81 wt.% of pure La10W2O21 phase and 19 wt.% of γ-La6W2O15. In this case, the maximum proton conductivity in wet air (1 × 10–3 S cm–1 at 700 °C) (Figure 11b) comparable to that of the known proton conductor La6–xWO12–δ (x = 0.6) was discovered [85]. Between two mirror composites with the γ-La6W2O15/La10W2O21 phases ratios of 82/18 and 19/81, the highest conductivity was demonstrated for the La10W2O21 based composite, with the difference in the protonic conductivity being in an order of magnitude (Figure 11b).
Figure 11. (a) Arrhenius plots of conductivity in dry (filled data points) and wet (open data points) air for γ-La6W2O15–based composites and pseudorhombohedral solid solutions, synthesized between 1450 and 1500 °C: (1) La10W2O21/γ-La6W2O15 composite (1450 °С); (2) La10W2O21/γ-La6W2O15/Nd6 composite (1500 °С); (3) La10W2O21/γ-La6W2O15/Nd8 composite (1500 °С); (4) La10W2O21/γ-La6W2O15/Nd10 composite (1500 °С); (5) La2Nd12W4O33 pseudorhombohedral solid solution (1500 °С); (6) Nd14W4O33 (1500 °С); (b) Arrhenius plots of conductivity in dry (1) and wet (2) air for the La10W2O21-based composite [152]. Reprinted from International Journal of Hydrogen Energy, Vol. 48, A.V. Shlyakhtina et al., Proton /oxygen ion conductivity ratio of Nd containing La10W2O21/γ-La6W2O15 tungstates, Pages No. 22671-22684, Copyright (2023), with permission from Elsevier.
Figure 11. (a) Arrhenius plots of conductivity in dry (filled data points) and wet (open data points) air for γ-La6W2O15–based composites and pseudorhombohedral solid solutions, synthesized between 1450 and 1500 °C: (1) La10W2O21/γ-La6W2O15 composite (1450 °С); (2) La10W2O21/γ-La6W2O15/Nd6 composite (1500 °С); (3) La10W2O21/γ-La6W2O15/Nd8 composite (1500 °С); (4) La10W2O21/γ-La6W2O15/Nd10 composite (1500 °С); (5) La2Nd12W4O33 pseudorhombohedral solid solution (1500 °С); (6) Nd14W4O33 (1500 °С); (b) Arrhenius plots of conductivity in dry (1) and wet (2) air for the La10W2O21-based composite [152]. Reprinted from International Journal of Hydrogen Energy, Vol. 48, A.V. Shlyakhtina et al., Proton /oxygen ion conductivity ratio of Nd containing La10W2O21/γ-La6W2O15 tungstates, Pages No. 22671-22684, Copyright (2023), with permission from Elsevier.
Preprints 115443 g011

2.4. Protonic Conductivity of Tetragonal Lanthanum Vanadates

LaVO4 compound have tetragonal (00-032-0504) and monoclinic (00-050-0367) crystal structures. In the new study [115] it is reported about the effect of partial substitution of La3+ in LaVO4 tetragonal ceramics with alkaline earth metals Ca2+, Sr2+ and Ba2+ (0.01 mol.) on the phase stability and the electrical properties. It was discovered that the acceptor doping led to increasing the ionic and protonic conductivity, and the substitution with Sr2+ gave the highest conductivity reaching ∼10–3 S cm–1 at 900 °C in wet O2 atmosphere. DFT calculations and molecular dynamics modeling ab initio demonstrated that protons generally form hydrogen bonds with the lattice oxygen near dopants and migrate via continuous process of hops and rotations between inter- and intratetrahedral groups VO4. V2O7 dimers form in the structure as well via common angles with neighboring VO4 isolated tetrahedral which allows to implement the ionic exchange via synergistic mechanism including destruction and reformation of V2O7 dimer.

2.5. Oxygen-Ion Conductivity in Triple Fluorite-Like Layers of Oxyhalides

Mixed-anion compounds, which contain two or more anions, are of great interest because of the wide variety of structures and properties [76,153,154,155].
Electron diffraction patterns of Bi2–xTexLuO4+x/2Cl (x = 0.1) also indicated the tetragonal Sillén phase. At any temperature, Bi1.9Te0.1LuO4.05Cl exhibited the highest σDC among these materials (e.g., 5.1 × 10–2 S cm–1 at 707 °C, Figure 12). The bulk conductivity of Bi1.9Te0.1LuO4.05Cl reaches 10–2 S cm–1 at 431 °C, which is much higher than of YSZ and La0.8Sr0.2Ga0.83Mg0.17O2.815 (LSGM) [156]. Thanks to the low activation energy, Bi1.9Te0.1LuO4.05Cl exhibits a high bulk conductivity of 1.5 × 10–3 S cm–1 even at a low temperature of 310 °C, which is 204 times higher than that of YSZ. The low activation energy is attributed to the interstitialcy oxide ion diffusion in the triple fluorite-like layer, as evidenced by neutron diffraction experiments (Rietveld and neutron scattering length density analyses), bond valence-based energy calculations, static DFT calculations, and ab initio molecular dynamics simulations. The electrical conductivity of Bi1.9Te0.1LuO4.05Cl is almost independent of the oxygen partial pressure from 10–18 to 10–4 atm at 431 °C, indicating the electrolyte domain. It is obvious that a significant electronic conductivity predominates while increasing the oxygen partial pressure for the samples of such a type, which definitely would prevent their application in SOFC [60], despite their phase and structural stability under a CO2 flow and ambient air at 400 °C. The oxide ion conduction due to the two-dimensional interstitialcy diffusion is considered to be common in Sillén oxyhalides with triple fluorite-like layers.

3. Cobaltite-Based Oxides with Layered Structure and Their Composites

3.1. Misfit Layered Oxides

Misfit layered oxides are characterized by a general formula AO–(MO)m–AO–CoO2 or (MmA2O2+m)qCoO2, where m = 1 or 2, A = Ca, Sr, Ba, etc., M = Co, Bi, etc. The layered structure of these compounds is comprised of repeating distorted rock salt (RS) type MmA2O2+m layers and CoO2 layers of a hexagonal (H) CdI2-like structure with incommensurate b-axes. The parameter q represents the incommensurability ratio (bRS/bH) that is typically close to 0.62 [79,80,157]. The example of misfit layered oxides is calcium cobaltite (CCO), which is composed of [Ca2CoO3±δ] layer sandwiched between two [CoO2] layers (Figure 13) [79,80,157,158,159,160]. The compound can be expressed with a general formula of [CoO2] [Ca2CoO3±δ]q (q = 0.614 – 0.69, typically 0.62) or Ca3Co4–xO9±δ (C349, CCO abbreviations). Both subsystems crystallize in a monoclinic crystal structure, where only the b lattice parameter varies.
Ca3Co4O9 is generally considered to be a thermoelectric material [80,161,162]. However, due to moderate values of the thermal expansion coefficient (TEC = 10–11 × 10–6 K–1 [163,164] providing excellent thermomechanical compatibility with the majority of solid state electrolytes, based on ZrO2, CeO2, LaGaO3 and BaCe(Zr)O3), mixed-conducting (e/O2–) and, in the case of proper doping, triple-conducting properties [165] and fast surface exchange kinetics (k* = 1.6 × 10–7 cm s–1 at 700 °C [160]), the interest in such compound as IT SOFC/SOE air electrode [157,159,165,166,167] and electrocatalyst [159,165,168] is growing.
Mixed ionic and electronic conduction observed in this material has been reported to originate from ionic conduction in oxygen deficient Ca2CoO3–δ′ layers and electronic conduction in the CoO2 layers [163]. The total conductivity of CCO, generally provided by the electronic component, ranges 101–102 S cm–1 at 700 °C depending on the synthesis technique and heat treatment mode [161,167,169,170,171,172,173,174,175]. It is related to high structural anisotropy of CCO material and low stability limit equal to 926 °C under air atmosphere, as can be seen in the phase diagram represented in Figure 14a. Therefore, CCO ceramics produced by conventional solid state sintering at 900 °C during 24 h possess a relatively low density (~ 50%) which leads to the formation of plate-like grains during the crystallization process (Figure 14b) and, as a result, decreased conductivity, while for highly dense samples obtained using spark plasma sintering (Figure 14c), conductivity reached 100 S cm–1.
Along with application of different synthesis techniques, a variety of approaches were employed to modify the electrical properties of CCO [176]. In the case of iso- and hetero-valence substitution on the Ca-site with alkaline earth [162,177,178] and alkali [176,179,180] metals, lanthanides [181,182], Y [183,184], Bi [185,186], Ag [178,187], Cu [188], the change in the conductivity was found to occur mainly due to modifying the carrier concentration and their mobility without fluctuating the band structure of the material. In the case of Bi-doping, a large increase in electrical conductivity was obtained due to the increased amount of Bi2O3 phase in the grain-boundary region [185]. Doping with basic monovalent Na+ ions in the Ca-site in CCO was found to generate a high density of extra charge carrier species with the increased Co oxidation state and facilitate both proton uptake and diffusion properties of CCO [180].
Figure 14. (a) Ca–Co–O phase diagram under air atmosphere [171]; a surface SEM image of CCO compact, fabricated using (b) a conventional solid-state sintering at 900 °C during 24 h [174] and (c) spark plasma sintering [175]. (a) Reprinted from Journal of the European Ceramic Society, Vol. 33, M.A. Madre et al., Preparation of high-performance Ca3Co4O9 thermoelectric ceramics produced by a new two-step method, Pages No. 1747-1754, Copyright (2013), with permission from Elsevier. (b) Reprinted from Journal of the Ceramics International, Vol. 41, G. Constantinescu et al., Effect of Na doping on the Ca3Co4O9 thermoelectric performance, Pages No. 10897-10903, Copyright (2015), with permission from Elsevier. (c) Reprinted from Journal of Solid State Chemistry, Vol. 313, Hohan Bae et al., Nonideal defect structure and high-temperature transport properties of misfit-layered cobalt oxide, Pages No. 123299, Copyright (2022), with permission from Elsevier.
Figure 14. (a) Ca–Co–O phase diagram under air atmosphere [171]; a surface SEM image of CCO compact, fabricated using (b) a conventional solid-state sintering at 900 °C during 24 h [174] and (c) spark plasma sintering [175]. (a) Reprinted from Journal of the European Ceramic Society, Vol. 33, M.A. Madre et al., Preparation of high-performance Ca3Co4O9 thermoelectric ceramics produced by a new two-step method, Pages No. 1747-1754, Copyright (2013), with permission from Elsevier. (b) Reprinted from Journal of the Ceramics International, Vol. 41, G. Constantinescu et al., Effect of Na doping on the Ca3Co4O9 thermoelectric performance, Pages No. 10897-10903, Copyright (2015), with permission from Elsevier. (c) Reprinted from Journal of Solid State Chemistry, Vol. 313, Hohan Bae et al., Nonideal defect structure and high-temperature transport properties of misfit-layered cobalt oxide, Pages No. 123299, Copyright (2022), with permission from Elsevier.
Preprints 115443 g014
The doping on the Co-site with transition-metal elements [189,190,191,192,193], Ga [194,195], Mo, W [191,196], Ti [162,197] was found to result in significant alterations to the band structure, particularly when doping in CoO2 layer. In examining the high-temperature transport properties of the Ca3Co4–xTixO9 system, it was observed that when the Ti content reached 0.3, Ti ions appeared in both CoO2 and Ca2CoO3 sublayers. The disturbance of the CoO2 layer resulted in a tendency for holes to jump to more distant low-energy sites instead of neighboring ones. This led to a change in the transport mechanism from thermally activated hopping to two-dimensional variable range hopping (2D-VRH) [197]. In contrast, temperature dependence of electrical conductivity of Ca3Co4–xMxO9+δ (M = Ni, Fe, Mn, x = 0, 0.2, 0.4, 0.6), considered to be substituted in the CoO2 layers, was found to demonstrate that the hole hopping conduction mechanism is dominant from room temperature up to 1000 K across the entire concentration range [192]. The exception was Ca3Co4–xCuxO9+δ (x = 0–0.3) samples, which exhibited a conductivity behavior similar to that of undoped CCO, displaying metallic-like behavior at low temperatures (below a transition temperature) and semiconducting behavior at high temperatures. It is noteworthy that the transition temperature increased with Cu doping. The authors proposed that Cu substitution occurred predominantly in the Ca2CoO3 layers, which resulted in an increase in CCO conductivity, while Ni, Fe, and Mn doping occurred in the CO2 layers, leading to a decrease in conductivity [192]. However, in the recent studies [189,198], Fe-doping was shown to occur in the Ca2CoO3 layers – at the Co1 site at low doping, and at Ca1 site at high doping [198]. It was demonstrated that Fe doping reduces the effective charge of O atoms in CCO, which results in the weakening of the co-valence bond and the formation of oxygen vacancies in CCO. This, in turn, increases the electrochemical activity of the related electrodes [189]. Furthermore, Fe doping enhances the hydration behavior, protonation, and proton migration (see Table 2).
Dual-doping has also been successfully employed to enhance the total electrical conductivity of CCO, including Na/Mo [164,199], Na/Cu and La/Cu [164,200], Gd/Cu [165], Y/Fe [201] systems.
Figure 15 summarizes temperature dependence of the total conductivity of some single and dual-doped systems. Among single-doped materials, the highest electrical conductivity values were obtained for Ca2.9Bi0.1Co4O9–δ (656 S cm–1 at 880 °C [202]), Ca0.85Na0.15Co4O9 (549 S cm–1 [180]) and Ca3Co3.9Cu0.1O9–δ (218.8 S cm–1 at 800 °C [191]) (Figure 15). Promising results were demonstrated for Ca2.9Ag0.4Co4O9 and Ca2.9Sr0.1Co3.92O9 (~145 S cm–1 at 700 °C [162,178]). However, it should be noted that Bi and Cu-doped systems demonstrate a sharp dependence of the total conductivity on temperature, which may be unfavorable from the application point. Stabilization of the conductivity across the entire range of the SOFC/SOEC application for the Cu-doped system was reached using dual-doping, for the Ca2.7La0.3Co3.8Cu0.2O9 system the total conductivity at 600 °C reached 187 S cm–1 [200].
An increase in the total conductivity of doped CCO resulted in an enhancement of the electrochemical performance of the related electrodes due to uniform current distribution across the electrode [165,180,182,184,188,202]. When employing undoped CCO as the electrode functional layer, it is advisable to utilize current collectors with higher conductivity due to insufficient conductivity of CCO with increasing electrode porosity. Current collectors can be fabricated from a noble metal (e.g., Au, Ag, with the exception of Pt because of its interaction with CCO) [159,203] or an oxide electronic conductor [158,159].
The oxygen diffusion in CCO is provided by oxygen vacancies localized in the middle of rock salt layer which can be generated since Co charge in this layer can be 2+, 3+ and 4+, and the formation of such vacancies is compensated by decreasing the content of electron holes in the CoO2 layer [159,160,168,175,204]:
O O × + 2 h 0.5   O 2 + V O + 2   e   .
The content of the vacancies depends on the oxygen partial pressure as well (Figure 16a). Moreover, the material’s stability also shows strong correlation with the temperature and oxygen partial pressure. Since in the SOFC mode, in the cells with an oxygen-ion conducting electrolyte, the oxygen partial pressure at the cathode reaches 10–3–10–4 atm, thus, the CCO electrode may perform without significant degradation in the temperature range of 600–700 °C (Figure 16b). Contrarily, the CCO electrode may successfully operates in a wide temperature range in the SOEC mode or in the SOFCs with a proton-conducting membrane, which was demonstrated in a number of studies [158,159,170,179,205,206].
Since Co cations’ charge ranges from 2+ to 4+, doping with 2+ cations such as, for instance, Cu2+ can increase the vacancy content, thus, along with an increase in the total conductivity, the ionic conductivity may also be increased [168,183]:
CuO C   u   Co + O O × + V O .
Unfortunately, the oxygen diffusion coefficient of CCO is not high enough (D* = 2.7 × 10–10 cm2 s–1 at 700 °С, Figure 17) [160,207]. Hence, the promising approach is using composite SOFC/SOE air electrodes comprised CCO and an oxide-ionic conductor such as Ce0.8Gd0.2O2–δ (CGO) [205] or Sm0.075Nd0.075Ce0.85O2–δ [208], as well as Ce0.8Pr0.2O2–δ [209], PrBaCo2O6–δ [167] or BaCe0.5Zr0.3Y0.1Yb0.1O3–δ [159], exhibiting partial electronic conductivity in air, or catalytically active additive PrOx [210]. An interesting approach was proposed by Guo et al. [211] to use a two-phase composite comprised CCO and La0.7Sr0.3CoO3 (LSC) with an acceptable level of the total conductivity (147 S cm–1 at 700 °C) and TEC (15.3 × 10–6 K–1) for the potential electrode application. Recently, Li et al. [212] proposed to use an interlayer based on La0.6Sr0.4FeO3–δ (LSF) of the CCO electrode and YSZ electrolyte interface, which allowed increasing twice hydrogen production efficiency in a reversible cell compared to that with CGO layer.
On the other hand, CCO possess a high oxygen surface reactivity (k* = 1.5×10–7 cm s–1 at 700 °C, Figure 17) apparently provided by the presence of weakly bound oxygen species on the surface which make available a fine performance in oxygen reduction/evolution reaction and, hence, compensate relatively low oxygen diffusion coefficient values and provide a good SOFC/SOE air electrode operation [159,160,165,206]. CCO-based materials demonstrate good transport properties in the presence of chemical potential gradient according to electrical conductivity relaxation data (Table 3) [189,207]. In the case of applying such a potential gradient, oxygen vacancies can easily be generated or occupied without significant surface exchange limitations. In the case of CCO-based electrode, this this can promote oxygen diffusion by rapid migration of oxygen vacancies from the electrolyte across the electrode – electrolyte interface [204].
CCO-based compounds can be used as catalysts for hydrogen and syngas production via sorption-enhanced steam reforming of ethanol (SESRE) and glycerol (SESRG) or their supports. Due to their transport features described above, CCO acts as an oxygen carrier is such a reaction. Furthermore, under the reaction conditions it can partially undergo a reversible phase change [37,38,213]:
Ca 3 Co 4 O 9   3 O 2 3 CaO + 4 Co ,
where CaO acts as a CO2 sorbent thus removing this product and shifting the reaction equilibrium, and metallic Co acts as a catalytically active component [37,38,213]. In particular, CCO may be involved in chemical looping methane combustion (CLMC, reaction (4), Figure 18) and then in the SESRG process (reaction (5)), and after that be regenerated (reaction (6)) [37,38,213]. According to the in situ XRD studies [213], the regeneration process (reaction (6)) proceeds step-wise via the formation of intermediate products such as CoO, Co3O4, CaCo2O4 and Ca2Co2O5 depending on the temperature.
Ca3Co4O9 (s) +1.5CH4 (g) = 3CaO(s) + 4Co(s) + 1.5CO2 (g) + 3H2O(g).
C3H8O3 (g) + 3H2O(g) + 3CaO = 7H2 (g) + 3CaCO3 (s)
3CaCO3 (s) + 4Co(s) + 3O2 (s) = Ca3Co4O9 (s) + 3CO2 (g)
There are three stages of SESRG process of CCO-based catalysts (Figure 19):
  • Pre-breakthrough: During this stage, a significant portion of the carbon dioxide generated by steam reforming and side water gas shift reactions is sorbed by the catalyst, thereby enhancing the production of hydrogen.
  • Breakthrough: At this stage, the sorption of carbon dioxide by CaO slows due to a decrease in the content of available sorption sites. Consequently, the sorption enhancement effect weakens.
  • Post-breakthrough: At this stage, the sorbent becomes fully saturated by carbon dioxide. The concentrations of products then reach their equilibrium values [37,38,213].
Figure 19. Typical SESRG product concentration depending on time [38,213].
Figure 19. Typical SESRG product concentration depending on time [38,213].
Preprints 115443 g019
The duration of these stages depends on the catalyst composition and the conditions.
Doping CCO with 1–2 wt. % of noble metals (such as Pt, Pd, Ru, Ir, Rh) increases glycerol conversion and hydrogen yield in pre- and post-breakthrough periods, and the best results have been demonstrated for 1 wt. % Pt-doped CCO (Figure 20) (glycerol conversion ~100 %, hydrogen yield ~90 %) [38]. In such catalysts, noble metals are exsolved as nanoparticles [37,38,213].

3.2. Ba Cobaltite Intergrowth Family

The other compounds based on layered cobaltite can be assumed while replacing the rock-salt layer in abovementioned misfit structure of layered cobaltites with a perovskite layer which results in another type of intergrowth structure. In such a structure, there are repeating n octahedral layers of perovskite [Ban+1ConO3n+3], CdI2-type layer 4[CoO2] and additional cobalt cations (4Co) in the interfacial layer resulting in a general formula of [Ban+1ConO3n+3] [Co8O8] for this intergrowth cobaltite family. The members of this family with n = 1, 2 and 3 are Ba2Co9O14 (often abbreviated as BCO), Ba3Co10O17 and Ba4Co11O20, respectively. In this structure, the Co cations are typically charged 2+ and 3+, e.g., Ba 2 Co 2 2 + Co 6 3 + O 14 and Ba 3 Co 2 2 + Co 8 3 + O 17 for n = 1 and 2, respectively. The other cation is Ba2+ which can be partially substituted by other alkaline earth metals and/or lanthanides. Figure 21 demonstrates the structure of Ba2Co9O14 [214,215,216].
Figure 20. Pre-breakthrough hydrogen yield in SESRG over reduced CCO-based catalysts: (a) CCO, (b) 1 wt. % Pt/CCO, (c) 2 wt. % Pt/CCO, (d) 2 wt. % Pd/CCO, (e) 2 wt. % Rh/CCO, (f) 2 wt. % Ru/CCO, (f) 2 wt. % Ir/CCO [38].
Figure 20. Pre-breakthrough hydrogen yield in SESRG over reduced CCO-based catalysts: (a) CCO, (b) 1 wt. % Pt/CCO, (c) 2 wt. % Pt/CCO, (d) 2 wt. % Pd/CCO, (e) 2 wt. % Rh/CCO, (f) 2 wt. % Ru/CCO, (f) 2 wt. % Ir/CCO [38].
Preprints 115443 g020
Figure 21. Intergrowth structure of the layered BCO containing the CdI2-based Co8O8 layers and perovskite Ba2CoO6 block [215]. Reprinted from Chemical Engineering Journal, Vol. 451, Minkyeong Jo et al., Layered barium cobaltite structure materials containing perovskite and CdI2-based layers for reversible solid oxide cells with exceptionally high performance, Pages No. 138954, Copyright (2023), with permission from Elsevier.
Figure 21. Intergrowth structure of the layered BCO containing the CdI2-based Co8O8 layers and perovskite Ba2CoO6 block [215]. Reprinted from Chemical Engineering Journal, Vol. 451, Minkyeong Jo et al., Layered barium cobaltite structure materials containing perovskite and CdI2-based layers for reversible solid oxide cells with exceptionally high performance, Pages No. 138954, Copyright (2023), with permission from Elsevier.
Preprints 115443 g021
Due to the mixed O2–+e type of conductivity, high total conductivity values (100–240 S cm–1 between 450 to 650 °C) and compatibility with major electrolytes types such as YSZ, GDC, LSGM, etc., reduced unwanted cation diffusion of the alkaline earth elements into the electrolyte, and reversible reducibility without structural destruction, BCO is considered as a promising material for SOFC/SOE oxygen electrodes (including protonic ceramic cells) [207,215,216,217,218] and electrocatalysts [215,219,220]. Moreover, BCO is capable of hydration accompanied by reduction of Co3+ cations and thus can be used as an electrocatalyst for hydrogen storage [220].
A further increase in the total conductivity can be achieved by doping BCO with various cations. E.g., for Nd0.1Ca0.1Ba1.8Co9O14, an increase in total conductivity compared to undoped BCO has been observed which has been attributed to the synergistic effects of co-doping of Nd3+ and Ca2+ in the Ba-site probably due to a combined effect of strong reactivity of Ca2+ with oxygen and high electronegativity of Nd3+ with a large structural parameter [215].
However, there are two major drawbacks of BCO. First, it has been reported that Ba2Co9O14 and Ba3Co10O17 are metastable at high temperatures (1000 °C and higher) and transform into 2H-perovskite-related oxides in the series (Ba8Co6O18)α(Ba8Co8O24)β [214]. On the other hand, BCO with Ba-site doped with Nd3+ and Ca2+ have demonstrated a stable performance with no observable degradation during 400 h of operation in SOFC and SOE modes [215]. At second, BCO possesses a low oxygen mobility according to the work [207] (Figure 22). On the other hand, the same work [207] states that it is impossible to derive accurate transport parameters since an oxidation process at the BCO surface is still not understood. Nevertheless, the promising approach is the use of composites based on BCO and solid electrolytes with a high oxygen mobility [216,217].

4. Other Materials

4.1. Magnetoplumbites and Related Structures

Magnetoplumbite-structured oxides with general formula AM12O19 (A are typically large cations such as Ca, Sr, Ba, Bb, Ln, etc.; M are typically small cations such as Fe, Al, Ga, etc.) are members of the magnetoplumbite-group materials, which also include diaoyudaoites (AM11O17), hibonites (AM12O19), nežilovites (AM12O19), plumboferrites (A2M11O19) and others [221,222,223,224,225,226,227]. Both A- and M-positions are often doped with various 2+, 3+ and 4+ cations. They are generally used as materials with specific magnetic properties, however, they are attracting attention as materials for SOFC/SOE electrodes [228] and catalysts for various processes such as methane partial oxidation, oxygen evolution reaction, etc. [229,230,231,232,233] due to their mixed ionic-electronic conductivity and moderate thermal expansion coefficient (~11 × 10–6 K–1 which is close to that of YSZ).
The AM12O19 magnetoplumbite structure is based on ten layers of approximately closest packing oxide anions perpendicular to the c-axis in the hexagonal unit cell. Two of these layers contain large A cations neighboring every fourth oxide anion. Small M cations are distributed over five interstitial sites: three with octahedral coordination, one with tetragonal one and one with trigonal-pyramidal one (coordination number 5). Among these layers, there are cubic blocks with spinel structure (S, [M6O8]2+) and hexagonal blocks (R, [AM6O11]2–), and these blocks repeat which can be written as RSR’S’, where R’ and S’ are rotated 180° around c-axis with respect to R and S blocks, respectively [221,225,228]. Figure 23 and Figure 24 demonstrate visualizations of the AM12O19 structure for aluminates and ferrites, respectively.
The mixed ionic-electronic character of the electrical conductivity of magnetoplumbites includes the contribution of metal cations, oxide anions, and electron transport. It is reported that the electrical conductivity of magnetoplumbites is anisotropic, and the conductivity along the hexagonal axis is about order of magnitude lower compared to that within the basal plate, and their temperature dependencies are similar [234,235]. Electronic conductivity can be both p-type and n-type and is assumed to be related to the presence of M2+ and M3+ cations (such as Fe2+/3+) that exchange with electrons [234,235,236,237]. Typical values of electronic conductivity vary with composition in the range of ~10–3–101 S cm–1 at 300 °C [238,239,240,241]. Aliovalent doping can lead to an increase in total conductivity due to the generation of additional charge carriers.
Metal cation diffusion also occurs in magnetoplumbites [237,239,240]. E.g., tracer diffusion coefficient values of Pb2+ are ~10–8 cm2 s–1 at 700 °C for PbFe12O19 and PbFe11O17.5, D* values of Sr2+ are ~10–10 cm2 s–1 at 900 °C for SrFe12O19 [239].
There is a lack of information on oxygen diffusion for magnetoplumbites. In different works [229,241], the oxygen diffusivity is characterized as from low to excellent. The oxygen-related defects may include oxygen vacancies and interstitial oxygen species [233,237]. It is also known that the oxygen diffusion is anisotropic: it is suppressed along the c-axis and prefers to take place within the mirror plane (Figure 23) [230,231,232,242]. The values of the oxygen tracer diffusion coefficient normal to the c-axis and along the c-axis measured for Ba hexaaluminate are 9.35 × 10–13 cm2 s–1 and 2.59 × 10–14 cm2 s–1 at 1400 °C, respectively [232]. It has also been reported that the oxygen diffusivity in the surface layer is faster compared to that within the bulk [233,243].
Magnetoplumbite-like compounds, including various hexaaluminates, are regarded as prospective catalysts or supports for many reactions of syngas production, generally for methane transformation reactions (partial oxidation, steam, dry and oxy-dry reforming) [39,230,231,233,244,245,246,247,248]. Such materials serve as oxygen carriers providing mobile oxygen species for participation in catalytic reactions.
Furthermore, under reducing conditions, they are capable of exsolving metal nanoparticles such as Ni0 and Fe0 which are reactive to methane oxidation into syngas. In particular, for MAl12–xNixO19-δ (M = alkaline earth or rare earth cation), Ni0 nanoparticles can be generated after the partial oxygen removal from the mirror planes and the reduction of Ni cations situated between the spinel block and the mirror planes followed by Ni0 species migration into the surface defect sites. Such a reducibility depends on the nature of the large cations present in the mirror plane. Due to this, doping with Ni leads to an increase in syngas yield and a prevention of the methane combustion, however, this may also lead to an increase or decrease in carbon formation depending on the catalyst composition and its synthesis conditions [231,244].
In the case of Fe-containing hexaaluminates, such as MAl12–xFexO19-δ (where M represents an alkaline earth or rare earth cation), Fe cations play important role in methane oxidation reactions. In particular, for Fe-doped Ba and La hexaaluminates, it was reported that some of Fe3+ cations occupying Al3+ sites suppress methane combustion and are reactive in methane oxidation into syngas. On the other hand, O6-Fe3+ (Oh) cations in LaFe3Al9O19 exhibit a high activity in methane combustion. Catalytically active Fe0 species generated on the surface contribute into increasing the syngas yield [130,233,245].
In the work authored by Z. Cheng et al. [233], the performance of CeO2/BaFe3Al9O19 catalysts in chemical looping dry reforming of methane was studied. Almost total conversion of methane and a high selectivity with respect to syngas were demonstrated. A proposed reaction mechanism is presented in Figure 25. Methane directly reacts with oxygen species in BaFe3Al9O19 which serves the dual function of oxygen carrier and catalyst. This results in a redistribution of oxygen within BaFe3Al9O19, which in turn affects the concentration of products. Initially, active surface oxygen species are consumed leading to the complete oxidation of methane:
CH4 (g) + 4MO(s) = 4M(s) + CO2 (g) + 2H2O(g).
Subsequently, the generation of oxygen vacancies on the surface facilitates the diffusion of less active bulk oxygen to the surface, thereby enabling its participation in the partial oxidation of methane into syngas:
CH4 (g) + MO(s) = M(s) + CO (g) + 2H2(g).
During the reaction, both CeO2 and BaFe3Al9O19 are gradually reduced to form CeFe1-xAlxO3 perovskite phase, which is conducive to the migration of lattice oxygen participating in partial oxidation of methane as well. The Ce3+ and Fe2+ cations with unsaturated coordination at the interface can function as active sites for methane adsorption and activation [233].
Figure 25. Reaction mechanism model for chemical looping dry reforming of methane over CeO2/BaFe3Al9O19 [233]. Reprinted from Fuel Processing Technology, Vol. 218, Zheng Cheng et al, Effect of calcination temperature on the performance of hexaaluminate supported CeO2 for chemical looping dry reforming, Pages No. 106873, Copyright (2021), with permission from Elsevier.
Figure 25. Reaction mechanism model for chemical looping dry reforming of methane over CeO2/BaFe3Al9O19 [233]. Reprinted from Fuel Processing Technology, Vol. 218, Zheng Cheng et al, Effect of calcination temperature on the performance of hexaaluminate supported CeO2 for chemical looping dry reforming, Pages No. 106873, Copyright (2021), with permission from Elsevier.
Preprints 115443 g025

4.2. Langasite Family

The langasite family of oxides with the general formula of A3BC3D2O12 is a large class of oxides generally recognized as piezoelectric crystals. In A3BC3D2O12, there are distinct cation sites:
  • A-site – dodecahedral sites;
  • B-site – octahedral site;
  • C-site – large tetrahedral sites;
  • D site – small tetrahedral sites.
B-site can be occupied by cations such as Mg2+, Sc3+, Ga3+, Ti4+, Zr4+, Ta5+, etc., C-site – by Zn2+, Fe3+, Ga3+, Ge4+, etc., D-site – by Al3+, Ga3+, Si4+, Ge4+, etc. [243,249,250,251,252].
The structure of langasite is complex and can be described as a mixed framework of two types of tetrahedra and one type of octahedra (Figure 26). The large holes of the framework (Thomson cubes) are occupied by large cations. The tetrahedra form layers that alternate along the c-axis with the layers consisting of octahedra and Thomson cubes. The structure of the langasite is controlled by the unique BC3O7 cluster which is linked by both the large dodecahedral A site and the small tetrahedral D site. Therefore, the cation size tolerance is much higher for both A and D sites. On the other hand, the cation size tolerance in both B and C sites is much smaller [249,250,251,252,253].
Langasite-type oxides can possess mixed ionic-electronic or pure ionic conductivity depending on composition and conditions. E.g., La3Ga5SiO14 possesses mixed ionic-electronic conductivity, however, at temperatures above 700 °C, the O2– type of conductivity predominates. Doping La3Ga5SiO14 with Sr reduces this temperature to 550 °C [243]. A similar effect has been demonstrated for La3Ta0.5Ga5.5O14 [254]. Langasites with a high dominant ionic conductivity such as Ge-doped La gallates (Figure 27) are considered as candidate materials for SOFC/SOE electrolytes [253,255,256].
The oxygen transport in langasites occurs via vacancy and interstitial mechanisms, or via a cooperative mechanism involving both regular and interstitial oxygen. For La3Ga6O14-based oxides, oxygen vacancy formation is energetically favorable, and the open structural framework of La3Ga5+xGe1–xO14–x/2 with structural and compositional flexibility is able to accommodate and transport oxygen vacancies. The oxygen vacancies favor the bridging O3 site between the octahedral and tetrahedral units. It was also demonstrated that disordered langasites possessed higher oxygen vacancy mobility compared to the ordered ones due to the oxygen vacancy ordering and trapping in deeper potential wells. The association of defects – oxygen vacancy and aliovalent dopant – can take place as well hindering oxygen vacancy transport [252,254,255,257]. Interstitial oxygen can be introduced into (Ga,Ge)2O8 structural unit of the La3Ga5–xGe1+xO14+0.5x langasite framework resulting in the structural deformation and disordering. Introduction of interstitial oxygen promotes chemical exchange between all oxide anions and supports the involvement of all oxide anions in the conduction mechanism, and its increases ionic motion as well. However, increasing its content is limited by requirement for large structure rearrangements to accumulate large amounts of (Ga,Ge)2O8 units, which acts as a trap for interstitial oxygen species and limits the ability to increase oxide ionic conductivity [253,256].
Acceptor doped langasites such as Ln3Ga5.06M0.94O14–δ (Ln = La, Nd; M = Si, Ti, Sn), La2.94Sr0.06Ga5SiO14–δ and La3Ga5+xSi1–xO14–δ (x = 0.06 and 0.12), La3Ga5+xTa1–xO14–δ (x = 0.525 and 0.6) demonstrate hydration ability with participation of oxygen vacancy and protonic contribution into the conductivity (σH up to ~10–4 S cm–1 at 900 °C). DFT calculations revealed anisotropy of proton transport in such langasites (Figure 28) [258,259,260].

4.3. Swedenborgites

Swedenborgites with the general formula of AT4OBO6, where A = alkali, alkaline earth, rare earth metals or cation vacancies, T = Be, Si, Co, Fe, Cu, Zn, Ga, etc., and B = Ba, Al, Sb, etc. possess a hexagonal structure, where A and B cations adopt 12- and 6-fold oxygen coordination, respectively, and the structure consists of corner-shared TO4 tetrahedra. In this structure, Kagomé layers of TO4 tetrahedra are connected by triangular layers (Figure 29) [261,262,263,264,265,266,267].
Co-containing swedenborgites RBaCo4–xMxO7 (R = Y, Ca, In, Lu, Yb, Tm, Er, Ho, Dy; M = Co, Zn, Fe, Al, Ga) such as YBaCo4O7+δ (YBC) and their composites with solid electrolytes such as GDC are promising materials for oxygen electrodes of intermediate-temperature SOFC/SOE (including proton ceramic cells) and oxygen separation membranes [264,265,266,267]. They possess very low TEC (8–11 × 10–6 K–1), which is very close to that of the BCZYYb electrolyte. However, phase decomposition at elevated temperatures of 700–800 °C has prevented their application. Doping with Zn increases the phase stability at high temperatures for RBa(Co,M)4O7 (R = Y, Ca, In; M = Zn, Fe, Al), while doping with Ca leads to deterioration of the phase stability. Employing a mixture of Y and In (50 % each) promotes phase stability and overcomes the phase decomposition problems [264,268,269]. In the Ga-doped YBaCo4–yGayO7+δ (y = 0.6–0.8) series, YBaCo3.2Ga0.8O7+δ showed a good stability in long-term studies. The Y-doped Y1–xInxBaCo3.3Ga0.7O7+δ (x = 0.1–0.9) series also remains stable at high temperatures, suggesting that the synergistic effect of In and Y could also maximize the stability at a certain Ga content [264,270].
YBC-based swedenborgites generally possess mixed O2–/e or triple H+/O2–/e conductivity [271,272,273,274]. Total conductivity varies on dopant nature and content, typical values being ~101 S cm–1 at 700 °C (Figure 30). Oxygen transference number is ~10–4 [275,276]. YBC-based swedenborgites are able to accumulate a large amounts of excess oxygen (7+δ = 7.0–9.0) a part of which is highly oxidative, which is provided by a good oxygen diffusion and surface exchange properties and the ability of Co cations to change their charge, making them promising as oxygen storage materials and oxygen reduction/evolution reaction electrocatalysts [265,266,267,271,275,276,277]. The Dchem and kchem values reported for Y0.8Er0.2BaCo3.2Ga0.8O7+δ are as high as ~10–3 cm2 s–1 and ~10–2 cm s–1 at 700 °C, respectively, which significantly exceed those of perovskite-based cathode materials [265]. Unfortunately, the data on the oxygen tracer diffusion coefficient for YBC-based compounds are lacking.

5. Conclusions and Perspectives

In this review, the selected types of materials for solid oxide fuel cells/electrolyzers, oxygen and hydrogen separation membranes and catalysts for fuel transformation into hydrogen and syngas beyond well-known perovskite- and fluorite-based oxides, spinels and other conventional and state-of-the-art materials are considered including various materials with non-cubic structure based on rare earth element tungstates, molybdates and vanadates, layered structured materials such as misfit layered oxides, other materials including oxides and solid solutions in magnetoplumbite and langasite families, and their composites. Literature data on such materials’ structural features, transport properties including electronic, oxide ionic and protonic conductivity, and their relationship as well as their use in the applications mentioned above are analyzed.
In design of materials for solid oxide fuel cells/electrolyzers, permselective membranes and efficient catalysts, the directions related to improving characteristics of perovskites and related materials (such as Ruddlesden – Popper phases, double perovskites), fluorites and related materials (such as pyrochlores, bixbyites, melitites), spinels, and composites based on them are still of current interests. Along with this, developing and studying new types of materials possessing ionic (O2– or/and H+) and mixed ionic-electronic (including triple H+/O2–/e) conductivity for these applications, and the search for promising materials in other, sometimes unexpected applications such as thermoelectric, barrier coating or piezoelectric materials are important topics here as well, since such materials can surprisingly exhibit fine functional properties, find these new applications, or even be considered as a new generation of SOFC/SOE, membrane and catalyst materials.
Unfortunately, some of these materials still are not chemically stable enough under operating conditions, which can be fixed by proper doping or other modification. Another problem is a lack in the detailed information on transport properties for some of such uncommon materials. Proper studying their electron, oxygen and proton transport features, and their relationship with composition and real/defect structure would enable more efficient performance of such materials as SOFC/SOE, permselective membrane or catalyst components.
Finally, theoretical approaches such as various simulations according to state-of-the-art mathematical models or methods related to using artificial intelligence would help in the search for new materials or in the modification of existing ones, in modeling their phase stability under different conditions, structural, electronic and ionic transport properties and other functional characteristics, and in predicting their performance as solid oxide fuel cell/electrolyzer components, hydrogen and oxygen separation membranes and catalysts for fuel transformation reactions.

Author Contributions

Conceptualization, V.S., N.E., A.S. and E.P.; writing—original draft preparation, N.E., A.S. and E.P.; writing—review and editing, V.S.; supervision, V.S. All authors have read and agreed to the published version of the manuscript.

Funding

Different parts of this work are supported by the Russian Federation Ministry of Science and Higher Education through the state research target for the Boreskov Institute of Catalysis, Siberian Branch, Russian Academy of Sciences (project FWUR-2024-0033), the Semenov Federal Research Center for Chemical Physics, Russian Academy of Sciences (RAS) (theme registration no. 122040500071-0: New Generation of Nanostructured Systems with Unique Functional Properties), and the Institute of High Temperature Electrochemistry, Ural Branch, Russian Academy of Sciences (theme registration no. 122020100324–3: Kinetics of discharge-ionization processes at the phase boundaries of solid oxide materials).

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Yousaf, M.; Lu, Y.; Akbar, M.; Lei, L.; Jing, S.; Tao, Y. Advances in Solid Oxide Fuel Cell Technologies: Lowering the Operating Temperatures Through Material Innovations. Mater. Today Energy 2024, 44, 101633. [CrossRef]
  2. Zamudio-García, J.; dos Santos-Gómez, L.; Losilla, E.R.; Marrero-López, D. Exploring Alkali Metal Doping in Solid Oxide Cells Materials: A Comprehensive Review. Chem. Eng. J. 2024, 493, 152832. [CrossRef]
  3. Singh, M.; Paydar, S.; Singh, A.K.; Singhal, R.; Singh, A.; Singh, M. Recent Advancement of Solid Oxide Fuel Cells Towards Semiconductor Membrane Fuel Cells. Energy Mater. 2024, 4, 400012. [CrossRef]
  4. Yuan, B.; Wang, N.; Tang, C.; Meng, L.; Du, L.; Su, Q.; Aoki, Y.; Ye, S. Advances and Challenges in High-Performance Cathodes for Protonic Solid Oxide Fuel Cells and Machine Learning-Guided Perspectives. Nano Energy 2024, 122, 109306. [CrossRef]
  5. Yadav, A.K.; Sinha, S.; Kumar, A. Advancements in Composite Cathodes for Intermediate-Temperature Solid Oxide Fuel Cells: A Comprehensive Review. Int. J. Hydrog. Energy 2024, 59, 1080–1093. [CrossRef]
  6. Chun, O.; Jamshaid, F.; Khan, M.Z.; Gohar, O.; Hussain, I.; Zhang, Y.; Zheng, K.; Saleem, M.; Motola, M.; Hanif, M.B. Advances in Low-Temperature Solid Oxide Fuel Cells: An Explanatory Review. J. Power Sources 2024, 610, 234719. [CrossRef]
  7. Afroze, S.; Reza, M.S.; Amin, M.R.; Taweekun, J.; Azad, A.K. Progress in Nanomaterials Fabrication and Their Prospects in Artificial Intelligence Towards Solid Oxide Fuel Cells: A Review. Int. J. Hydrog. Energy 2024, 52, 216–247. [CrossRef]
  8. Baratov, S.; Filonova, E.; Ivanova, A.; Hanif, M.B.; Irshad, M.; Khan, M.Z.; Motola, M.; Rauf, S.; Medvedev, D. Current and Further Trajectories in Designing Functional Materials for Solid Oxide Electrochemical Cells: A Review of Other Reviews. J. Energy Chem. 2024, 94, 302–331. [CrossRef]
  9. Ahmed, N.; Devi, S.; Dar, M.A.; Ibrahim, S.K.M.; Sharma, A.; Sharma, N.; Paul, S.; Ahamed, S.R. Anode Material for Solid Oxide Fuel Cell: A Review. Indian J. Phys. 2024, 98, 877–888. [CrossRef]
  10. Fan, K.; Yu, M.; Lei, J.; Mu, S. Advanced Preparation Methods for Ceramic Membrane Materials in Electrochemical Applications. Crystals 2024, 14, 623. [CrossRef]
  11. Kumar, R.; Singh, R. Dutta, S. Review and Outlook of Hydrogen Production through Catalytic Processes. Energy Fuels 2024, 38, 2601–2629. [CrossRef]
  12. Jiao, Y.; Dai, J.; Fan, Z.; Cheng, J.; Zheng, G.; Grema, L.; Zhong, J.; Li, H.-F.; Wang, D. Overview of High-Entropy Oxide Ceramics. Mater. Today 2024, in press. [CrossRef]
  13. Chen, G.; Widenmeyer, M.; Yu, X.; Han, N.; Tan, X.; Homm, G.; Liu, S.; Weidenkaff, A. Perspectives on Achievements and Challenges of Oxygen Transport Dual-Functional Membrane Reactors. J. Am. Ceram. Soc. 2024, 107, 1490–1504. [CrossRef]
  14. Osman, A.I.; Nasr, M.; Farghali, M.; Bakr, S.S.; Eltaweil, A.S.; Rashwan, A.K., Abd El-Monaem, E.M. Machine Learning for Membrane Design in Energy Production, Gas Separation, and Water Treatment: A Review. Environ. Chem. Lett. 2024, 22, 505–560. [CrossRef]
  15. Qu, L.; Papaioannou, E.I. Development of Mixed Ionic and Electronic Conducting Materials for Gas Separation Membranes: A Critical Overview. Chem. Eng. J. 2024, 496, 153791. [CrossRef]
  16. Singh, R.; Prasad, B.; Ahn, Y.-H. Recent Developments in Gas Separation Membranes Enhancing the Performance of Oxygen and Nitrogen Separation: A Comprehensive Review. Gas Sci. Eng. 2024, 123, 205256. [CrossRef]
  17. Pradhan, P.; Rathod, A.P.; Rai, S.B.; Mohapatra, S.S. An Overview of Research Progress on Ceramic-Based Membranes. Mater. Today Proc. 2023, in press. [CrossRef]
  18. Naeem, A.; Saeed, B.; AlMohamadi, H.; Lee, M.; Amjad Gilani, M.; Nawaz, R.; Khan, A.L.; Yasin, M. Sustainable and Green Membranes for Chemical Separations: A Review. Sep. Purif. Technol. 2024, 336, 126271. [CrossRef]
  19. Boscherini, M.; Storione, A.; Minelli, M.; Miccio, F.; Doghieri, F. New Perspectives on Catalytic Hydrogen Production by the Reforming, Partial Oxidation and Decomposition of Methane and Biogas. Energies 2023, 16, 6375. [CrossRef]
  20. Meloni, E.; Martino, M.; Iervolino, G.; Ruocco, C.; Renda, S.; Festa, G.; Palma, V. The Route from Green H2 Production through Bioethanol Reforming to CO2 Catalytic Conversion: A Review. Energies 2022, 15, 2383. [CrossRef]
  21. Dey, A.; Gogate, P.R.; Gote, Y.M. A Review on Ultrasound Assisted Synthesis of Metal Oxide and Doped Metal Oxide Nanocatalysts and Subsequent Application as Photocatalyst for Dye Degradation. Environ. Qual. Manage. 2024, 33, 139–163. [CrossRef]
  22. Zhang, M.; Guan, Z.; Qiao, Y.; Zhou, S.; Chen, G.; Guo, R.; Pan, W.; Wu, J.; Li, F.; Ren, J. The Impact of Catalyst Structure and Morphology on the Catalytic Performance in NH3-SCR Reaction: A Review. Fuel 2024, 361, 130541. [CrossRef]
  23. Ebrahimzadeh Sarvestani, M.; Norouzi, O.; Di Maria, F.; Dutta, A. From Catalyst Development to Reactor Design: A Comprehensive Review of Methanol Synthesis Techniques. Energy Convers. Manage. 2024, 302, 118070. [CrossRef]
  24. Wang, P.; Shi, R.; Zhao, J.; Zhang, T. Photodriven Methane Conversion on Transition Metal Oxide Catalyst: Recent Progress and Prospects. Adv. Sci. 2024, 11, 2305471. [CrossRef]
  25. Long, F.; Ghani, D.; Huang, R.; Zhao, C. Versatile Electrode Materials Applied in the Electrochemical Advanced Oxidation Processes for Wastewater Treatment: A Systematic Review. Sep. Purif. Technol. 2025, 354, 128725. [CrossRef]
  26. Yao, X.; Cheng, Q.; Attada, Y.; Ould-Chikh, S.; Ramírez, A.; Bai, X.; Mohamed, H.O.; Li, Q.; Shterk, G.; Zheng, L.; Gascon, J.; Han, Y.; Bakr, O.M.; Castaño, P. Atypical Stability of exsolved Ni-Fe Alloy Nanoparticles on Double Layered Perovskite for CO2 Dry Reforming of Methane. Appl. Catal. B.: Environ. 2023, 328, 122479. [CrossRef]
  27. Georgiadis, A.G.; Charisiou, N.D.; Goula, M.A. A Mini-Review on Lanthanum–Nickel-Based Perovskite-Derived Catalysts for Hydrogen Production via the Dry Reforming of Methane (DRM). Catalysts 2023, 13, 1357. [CrossRef]
  28. Jawad, A.; Rezaei, F.; Rownaghi, A.A. Highly Efficient Pt/Mo-Fe/Ni-Based Al2O3-CeO2 Catalysts for Dry Reforming of Methane. Catal. Today 2020, 350, 80–90. [CrossRef]
  29. Kwon, O.; Huang, R.; Cao, T.; Vohs, J.M.; Gorte, R.J. Dry Reforming of Methane over Ni Supported LaMnO3 Thin Films. Catal. Today 2021, 382, 142–147. [CrossRef]
  30. Wang, E.; Li, L.; Cui, C.; Costa, P.Da.; Hu, C. The Effect of Adsorbed Oxygen Species on Carbon-Resistance of Ni-Zr Catalyst Modified by Al and Mn for Dry Reforming of Methane. Catal. Today 2022, 384–386, 257–264. [CrossRef]
  31. Miri, S.S.; Meshkani, F.; Rastegarpanah, A.; Rezaei, M. Influence of Fe, La, Zr, Ce, and Ca on the Catalytic Performance and Coke Formation in Dry Reforming of Methane over Ni/MgO.Al2O3 Catalyst. Chem. Eng. Sci. 2022, 250, 116956. [CrossRef]
  32. Shafiqah, M.-N.N.; Siang, T.J.; Senthil Kumar, P.; Ahmad, Z.; Jalil, A.A.; Bahari, M.B.; Le, Q.V.; Xiao, L.; Mofijur, M.; Xia, C.; Ahmed, S.F.; Vo, D.-V.N. Advanced catalysts and Effect of Operating Parameters in Ethanol Dry Reforming for Hydrogen Generation. A Review. Environ. Chem. Lett. 2022, 20, 1695–1718. [CrossRef]
  33. Ogo, S.; Sekine, Y. Recent Progress in Ethanol Steam Reforming using Non-Noble Transition Metal Catalysts: A Review. Fuel 2020, 199, 106238. [CrossRef]
  34. Setiabudi, H.D.; Aziz, M.A.A.; Abdullah, S.; Teh, L.P.; Jusoh, R. Hydrogen Production from Catalytic Steam Reforming of Biomass Pyrolysis Oil or Bio-Oil Derivatives: A Review. Int. J. Hydrog. Energy 2020, 45, 18376–18397. [CrossRef]
  35. Kwon, O.; Foucher, A.C.; Huang, R.; Stach, E.A.; Vohs, J.M.; Gorte, R.J. Evidence for Redispersion of Ni on LaMnO3 Films Following High-Temperature Oxidation. J. Catal. 2022, 407, 213–220. [CrossRef]
  36. Du, H.; Luo, H.; Jiang, M.; Yan, X.; Jiang, F.; Chen, H. A Review of Activating Lattice Oxygen of Metal Oxides for Catalytic Reactions: Reaction Mechanisms, Modulation Strategies of Activity and Their Practical Applications. Appl. Catal. A.: Gen. 2023, 664, 119348. [CrossRef]
  37. Xu, Y.; Wu, M.; Yang, X.; Sun, S.; Li, Qi.; Zhang, Y.; Wu, C.; Przekop, R.E.; Romańczuk-Ruszuk, E.; Pakuła, D.; Zhou, H. Recent Advances and Prospects in High Purity H2 Production from Sorption Enhanced Reforming of Bio-Ethanol and Bio-Glycerol as Carbon Negative Processes: A Review. Carbon Capture Sci. Technol. 2023, 8, 100129. [CrossRef]
  38. Li, H.; Dang, X.; Li, Y.; Yang, G.; Cao, Y.; Wang, H.; Peng, F.; Yu, H. Pt–Calcium Cobaltate Enables Sorption-Enhanced Steam Reforming of Glycerol Coupled with Chemical-Looping CH4 Combustion. AICHe J. 2021, 67, e17383. [CrossRef]
  39. Zhang, Z.; Zhang, Y.; Liu, L. Role and Mechanism of Calcium-Based Catalysts for Methane Dry Reforming: A Review. Fuel 2024, 355, 129329. [CrossRef]
  40. Starostin, G.; Volkov, A.N.; Kalyakin, A.S.; Medvedev, D.A. High-Temperature Gas Sensors Based on Proton-Conducting Ceramic Oxides. A Brief Review. Ceram. Int. 2024, in press. [CrossRef]
  41. Morulane, K.L.; Swart, H.C.; Motaung D.E. A Review on Topical Advancement and Challenges of Indium Oxide Based Gas Sensors: Future Outlooks. J. Environ. Chem. Eng. 2024, 12, 112144. [CrossRef]
  42. Kanan, S.; Obeideen, K.; Moyet, M.; Abed, H.; Khan, D.; Shabnam, A.; El-Sayed, A.; Arooj, M.; Mohamed, A.A. Recent Advances on Metal Oxide Based Sensors for Environmental Gas Pollutants Detection. Crit. Rev. Anal. Chem. 2024, Latest Articles. [CrossRef]
  43. Feng, X.; Cheng, R.; Yin, L.; Wen, Y.; Jiang, J.; He, J. Two-Dimensional Oxide Crystals for Device Applications: Challenges and Opportunities. Adv. Mater. 2024, 36, 2304708. [CrossRef]
  44. Khan, A.; Muhammad, R.; Farhan, M.; Shafi, H.Z.; Shah, A.U.; Hussain, F.; Han, D.; Wang, D. Lan+1NinO3n+1 (n = 1, 2, 3) RP-Phases for EMI Shielding and ESD Applications. Ceram. Int. 2024, 50, 33392–33399. [CrossRef]
  45. Bosu, S.; Rajamohan, N. Recent Advancements in Hydrogen Storage - Comparative Review on Methods, Operating Conditions and Challenges. Int. J. Hydrog. Energy 2024, 52, 352–370. [CrossRef]
  46. Wang, X.; Zhong, Y.; Hu, Q. A Review of Al2O3-Based Eutectic Ceramics for High-Temperature Structural Materials. J. Mater. Sci. Technol. 2024, in press. [CrossRef]
  47. Kuzieva, M.A. The Use of Ceramic Materials in Orthopedic Dentistry. (Literature Review). Tadqiqotlar 2024, 31, 75–85.
  48. Nisar, J.; Kaur, G.; Giddey, S.; Bhargava, S.; Jones, L. Cathode Materials for Intermediate Temperature Solid Oxide Fuel Cells. Preprints 2024, 2024050747.
  49. Nechache, A.; Hody, S. Alternative and Innovative Solid Oxide Electrolysis Cell Materials: A Short Review. Renew. Sustain. Energy Rev. 2021, 149, 111322. [CrossRef]
  50. Shah, N.; Xu, X.; Love, J.; Wang, H.; Zhu, Z.; Ge, L. Mitigating Thermal Expansion Effects in Solid Oxide Fuel Cell Cathodes: A Critical Review. J. Power Sources 2024, 599, 234211. [CrossRef]
  51. Khan, M.Z.; Song, R.-H.; Taqi Mehran, M.; Lee, S.-B.; Lim, T.-H. Controlling Cation Migration and Inter-Diffusion Across Cathode/Interlayer/Electrolyte Interfaces of Solid Oxide Fuel Cells: A Review. Ceram. Int. 2021, 47, 5839–5869. [CrossRef]
  52. Salehabadi, A. Rare-Earth Based Tungstates Ceramic Materials: Recent Advancements and Technologies. In Advanced Rare Earth-Based Ceramic Nanomaterials; Zinatloo-Ajabshir, S., Ed.; Elsevier, 2022; pp. 175–204.
  53. Tsvetkov, D.S.; Sereda, V.V.; Malyshkin, D.A.; Ivanov I.L.; Zuev, A.Yu. Chemical Lattice Strain in Nonstoichiometric Oxides: An Overview. J. Mater. Chem. A 2022, 10, 6351–6375. [CrossRef]
  54. Ma, Z.; Ye, Q.; Ye, H.; Dong, F.; Ni M.; Lin, Z. CO2-Tolerant Perovskite Cathodes for Enhanced Solid Oxide Fuel Cells: Advancements, Challenges, and Strategic Perspectives. J. Mater. Chem. A 2024, 12, 19606–19626. [CrossRef]
  55. Samreen, A.; Ali, M.S.; Huzaifa, M.; Ali, N.; Hassan, B.; Ullah, F.; Ali, S.; Arifin, N.A. Advancements in Perovskite-Based Cathode Materials for Solid Oxide Fuel Cells: A Comprehensive Review. Chem. Rec. 2024, 24, e202300247. [CrossRef]
  56. Omeiza, L.A.; Kabyshev, A.; Bekmyrza, K.; Kuterbekov, K.A.; Kubenova, M.; Zhumadilova, Z.A.; Subramanian, Y.; Ali, M.; Aidarbekov, N.; Azad, A.K. Constraints in Sustainable Electrode Materials Development for Solid Oxide Fuel Cell: A Brief Review. Mater. Sci. Energy Technol. 2025, 8, 32–43. [CrossRef]
  57. Medvedev, D.A. Current Drawbacks of Proton-Conducting Ceramic Materials: How to Overcome Them for Real Electrochemical Purposes. Curr. Opin. Green Sustain. Chem. 2021, 32, 100549. [CrossRef]
  58. Tarutin, A.P.; Filonova, E.A.; Ricote, S.; Medvedev, D.A.; Shao, Z. Chemical Design of Oxygen Electrodes for Solid Oxide Electrochemical Cells: A Guide. Sustain. Energy Technol. Assess. 2023, 57, 103185. [CrossRef]
  59. Chroneos, A.; Goulatis, I.L.; Solovjov, A.; Vovk, R.V. The Evolution of Solid Oxide Fuel Cell Materials. Appl. Sci. 2024, 14, 69. [CrossRef]
  60. Filonova, E.; Medvedev, D. Recent Progress in the Design, Characterisation and Application of LaAlO3- and LaGaO3-Based Solid Oxide Fuel Cell Electrolytes. Nanomaterials 2022, 12, 1991. [CrossRef]
  61. Sadykov, V.; Pikalova, E.; Sadovskaya, E.; Shlyakhtina, A.; Filonova, E.; Eremeev, N. Design of Mixed Ionic-Electronic Materials for Permselective Membranes and Solid Oxide Fuel Cells Based on Their Oxygen and Hydrogen Mobility. Membranes 2023, 13, 698. [CrossRef]
  62. Sadykov, V.; Eremeev, N.; Sadovskaya, E.; Bespalko, Yu.; Simonov, M.; Arapova, M.; Smal, E. Nanomaterials with Oxygen Mobility for Catalysts of Biofuels Transformation into Syngas, SOFC and Oxygen/Hydrogen Separation Membranes: Design and Performance. Catal. Today 2023, 423, 113936. [CrossRef]
  63. Sadykov, V.; Simonov, M.; Eremeev, N.; Mezentseva, N. Modern Trends in Design of Catalysts for Transformation of Biofuels into Syngas and Hydrogen: From Fundamental Bases to Performance in Real Feeds. Energies 2021, 14, 6334. [CrossRef]
  64. Fop, S. Solid Oxide Proton Conductors Beyond Perovskites. J. Mater. Chem. A 2021, 9, 18836–18856.
  65. Morales-Zapata, M.A.; Larrea, A.; Laguna-Bercero, M.A. Lanthanide Nickelates for Their Application on Solid Oxide Cells. Electrochim. Acta 2023, 444, 141970. [CrossRef]
  66. Pikalova, E.Yu.; Guseva, E.M.; Filonova, E.A. Short Review on Recent Studies and Prospects of Application of Rare-Earth-Doped La2NiO4+δ as Air Electrodes for Solid Oxide Electrochemical Cells. EM&T. 2023, 2, 20232025.
  67. Xie, M.; Cai, C.; Duan, X.; Xue, K.; Yang, H.; An, S. Review on Fe-Based Double Perovskite Cathode Materials for Solid Oxide Fuel Cells. Energy Mater. 2024, 4, 400007. [CrossRef]
  68. Klyndyuk, A.I.; Chizhova, E.A.; Kharytonau, D.S.; Medvedev, D.A. Layered Oxygen-Deficient Double Perovskites as Promising Cathode Materials for Solid Oxide Fuel Cells. Materials 2022, 15, 141. [CrossRef]
  69. Kang, K.T. Review of the Fabrication of SrFeO2.5 Thin Film and Its Functionalities. New Phys.: Sae Mulli 2023, 73, 210–221. [CrossRef]
  70. Erpalov, M.V.; Tarutin, A.P.; Danilov, N.A.; Osinkin, D.A.; Medvedev, D.A. Chemistry and Electrochemistry of CeO2-Based Interlayers: Prolonging the Lifetime of Solid Oxide Fuel and Electrolysis Cells. Russ. Chem. Rev. 2023, 92, RCR5097. [CrossRef]
  71. Das, S.; Bhaskar, R.; Narayanan, K.B. Multifunctional Applications of Gadolinium-Doped Cerium Oxide (Ce1–xGdxO2–) Ceramics: A Review. J. Rare Earths 2024, in press. [CrossRef]
  72. Fukina, D.G.; Suleimanov, E.V. Structural Type of α-Pyrochlore Oxides. In Pyrochlore Oxides. Structure, Properties, and Potential in Photocatalytic Applications; Fukina, D.G., Belousov, A.S., Suleimanov, E.V., Eds.; Springer Cham, 2024; pp. 1–36. [CrossRef]
  73. Yang, X.; Fernández-Carrión, A.J.; Kuang, X. Oxide Ion-Conducting Materials Containing Tetrahedral Moieties: Structures and Conduction Mechanisms. Chem. Rev. 2023, 123, 9356–9396. [CrossRef]
  74. Sun, S.; Tang, Q.; Zhang, K.; Wen, Y.; Billings, A.; Huang, K. A Focused Review on Structures and Ionic Conduction Mechanisms in Inorganic Solid-State Proton and Hydride Anion Conductors. Mater. Adv. 2023, 4, 389–407.
  75. Chong, S.; Riley, B.J.; Lu, X.; Du, J.; Mahadevan, T.; Hegde, V. Synthesis and Properties of Anhydrous Rare-Earth Phosphates, Monazite and Xenotime: A Review. RSC Adv. 2024, 14, 18978–19000. [CrossRef]
  76. Cheng, H. Rare Earth Tungstate: One Competitive Proton Conducting Material Used for Hydrogen Separation: A Review. Separations 2023, 10, 317. [CrossRef]
  77. Granadeiro, C.M.; Julião, D.; Ribeiro, S.O.; Cunha-Silva, L.; Balula, S.S. Recent Advances in Lanthanide-Coordinated Polyoxometalates: From Structural Overview to Functional Materials. Coord. Chem. Rev. 2023, 476, 214914. [CrossRef]
  78. Pautonnier, A.; Coste, S.; Barré, M.; Lacorre, P. Higher Lanthanum Molybdates: Structures, Crystal Chemistry and Properties. Prog. Solid State Chem. 2023, 69, 100382. [CrossRef]
  79. Ng, N.; McQueen, T.M. Misfit Layered Compounds: Unique, Tunable Heterostructured Materials with Untapped Properties. APL Mater. 2022, 10, 100901. [CrossRef]
  80. Aliev, S.B.; Tenne, R. Quaternary Misfit Compounds—A Concise Review. Crystals 2020, 10, 468. [CrossRef]
  81. Tarasova, N.; Hanif, M.B.; Janjua, N.K.; Anwar, S.; Motola, M.; Medvedev, D. Fluorine-Insertion in Solid Oxide Materials for Improving Their Ionic Transport and Stability. A Brief Review. Int. J. Hydrog. Energy 2024, 50, 104–123. [CrossRef]
  82. Butz, B.; Kruse, P.; Störmer, H.; Gerthsen, D.; Müller, A.; Weber, A.; Ivers-Tiffée, E. Correlation Between Microstructure and Degradation in Conductivity for Cubic Y2O3-Doped ZrO2. Solid State Ionics 2006, 177, 3275–3284. [CrossRef]
  83. Lacorre, P.; Goutenoire, F.; Bohnke, O.; Retoux, R.; Laligant, Y. Designing fast oxide-ion conductors based on La2Mo2O9. Nature 2000, 404, 856–858. [CrossRef]
  84. Shimura, T.; Fujimoto, S.; Iwahara, H. Proton Conduction in Non-Perovskite-Type Oxides at Elevated Temperatures. Solid State Ionics 2001, 143, 117–123. [CrossRef]
  85. Magrasó, A.; Polfus, J.M.; Frontera, C.; Canales-Vázquez, J.; Kalland, L.-E.; Hervoches, C.H.; Erdal, S.; Hancke, R.; Islam, M.S.; Norby, T.; Haugsrud, R. Complete Structural Model for Lanthanum Tungstate: A Chemically Stable High Temperature Proton Conductor by Means of Intrinsic Defects. J. Mater. Chem. 2012, 22, 1762–1764. [CrossRef]
  86. Magrasó, A.; Haugsrud, R. Effects of the La/W Ratio and Doping on the Structure, Defect Structure, Stability and Functional Properties of Proton-Conducting Lanthanum Tungstate La28−xW4+xO54+δ. A Review. J. Mater. Chem. A 2014, 2, 12630–12641. [CrossRef]
  87. Escolastico, S.; Seeger, J.; Roitsch, S.; Ivanova, M.; Meulenberg, W.A.; Serra, J.M. Enhanced H2 Separation through Mixed Pro-ton-Electron Conducting Membranes Based on La5.5W0.8M0.2O11.25−δ. ChemSusChem 2013, 6, 1523–1532. [CrossRef]
  88. Magrasó, A. Transport Number Measurements and Fuel Cell Testing of Undoped and Mo-Substituted Lanthanum Tungstate. J. Power Sources 2013, 240, 583–588. [CrossRef]
  89. Polfus, J.M.; Li, Z.; Xing, W.; Sunding, M.F.; Walmsley, J.C.; Fontaine, M.-L.; Henriksen, P.P.; Bredesen, R. Chemical Stability and H2 Flux Degradation of Cercer Membranes Based on Lanthanum Tungstate and Lanthanum Chromite. J. Membr. Sci. 2016, 503, 42–47. [CrossRef]
  90. Goutenoire, F.; Isnard, O.; Retoux, R.; Lacorre, P. Crystal Structure of La2Mo2O9, a New Fast Oxide−Ion Conductor. Chem. Mater. 2000, 12, 2575–2580. [CrossRef]
  91. Alekseeva, O.A.; Verin, I.A.; Sorokina, N.I.; Kharitonova, E.P.; Voronkova, V.I. Structure Properties of Antimony-Doped Lanthanum Molybdate La2Mo2O9. Crystallogr. Rep. 2011, 56, 435–442. [CrossRef]
  92. Alekseeva, O.A.; Sorokina, N.I.; Verin, I.A.; Voronkova, V.I.; Krasil’nikova A.E. Crystal Structure of the Metastable Cubic B1 Phase of the La2Mo2O9 Single Crystal. Crystallogr. Rep. 2009, 54, 19–24. [CrossRef]
  93. Orlova, E.I.; Morkhova, Ye.A.; Egorova, A.V.; Kharitonova, E.P.; Lyskov, N.V.; Voronkova, V.I.; Kabanov, A.A.; Veligzhanin, A.A.; Kabanova, N.A. Mechanism of Conductivity in the Rare Earth Layered Ln2MoO6 (Ln = La, Pr, and Nd) Oxymolybdates: Theoretical and Experimental Investigations. J Phys. Chem. C 2022, 126, 9623−9633. [CrossRef]
  94. Morkhova, Y.A.; Orlova, E.I.; Kabanov, A.A.; Sorokin, T.A.; Egorova, A.V.; Gilev, A.R.; Kharitonova, E.P.; Lyskov, N.V.; Voronkova, V.I.; Kabanova, N.A. Comprehensive Study of Conductivity in the Series of Monoclinic Oxymolybdates: Ln2MoO6 (Ln = Sm, Gd, Dy). Solid State Ionics 2023, 400, 116337. [CrossRef]
  95. Li, Q.; Thangadurai, V. Novel Nd2WO6-Type Sm2−xAxM1−yByO6−δ (A = Ca, Sr; M= Mo, W; B = Ce, Ni) Mixed Conductors. J. Power Sources 2011, 196, 169–178. [CrossRef]
  96. Sun, C.W.; Stimming, U. Recent Anode Advances in Solid Oxide Fuel Cells. J. Power Sources 2007, 171, 247–260. [CrossRef]
  97. Chambrier, M.-H.; Kodjikian, S.; Ibberson, R.M.; Goutenoire, F. Ab-Initio Structure Determination of β-La2WO6. J. Solid State Chem. 2009, 182, 209–214. [CrossRef]
  98. Allix, M.; Chambrier, M.-H.; Véron, E.; Porcher, F.; Suchomel, M.; Goutenoire, F. Synthesis and Structure Determination of the High Temperature Form of La2WO6. Cryst. Growth Des. 2011, 11, 5105–5112. [CrossRef]
  99. Chambrier, M.-H.; Le Bail, A.; Kodjikian, S.; Suard, E.; Goutenoire, F. Structure Determination of La18W10O57. Inorg. Chem. 2009, 48, 6566–6572. [CrossRef]
  100. Yanovskii, V.K.; Voronkova, V.I. Polytypism of La2WO6 Crystals. Kristallografiya 1981, 26, 604–606.
  101. Chambrier, M.-H.; Ibberson, R.M.; Goutenoire, F. Structure Determination of α-La6W2O15. J. Solid State Chem. 2010, 183, 1297–1302. [CrossRef]
  102. Chambrier, M.-H.; Le Bail, A.; Giovannelli, F.; Redjaïmia, A.; Florian, P.; Massiot, D.; Suard, E.; Goutenoire, F. La10W2O21: An Anion-Deficient Fluorite-Related Superstructure with Oxide Ion Conduction. Inorg. Chem. 2014, 53, 147–159. [CrossRef]
  103. Kovalevsky, A.V.; Kharton, V.V.; Naumovich, E.N. Oxygen Ion Conductivity of Hexagonal La2W1.25O6.75. Mater. Lett. 1999, 38, 300–304. [CrossRef]
  104. Vigen, C.K.; Pan, J.; Haugsrud, R. Defects and Transport in Acceptor Doped La2WO6 and Nd1.2Lu0.8WO6. ECS J. Solid State Sci. Technol. 2013, 2, N243–N248. [CrossRef]
  105. Shlyakhtina, A.V.; Lyskov, N.V.; Kolbanev, I.V.; Vorob’eva, G.A.; Shchegolikhin, A.N.; Voronkova, V.I. Specific Features of Phase Formation and Properties of Compounds La2W1+xO6+3x (x ~ 0; 0.11–0.22). Russ. J. Electrochem. 2023, 59, 60–69. [CrossRef]
  106. Ivanova, M.E.; Seeger, J.; Serra, J.M.; Solis, C.; Meulenberg, W.A.; Roitsch, S.; Buchkremer, H.P. Influence of the La6W2O15 Phase on the Properties and Integrity of La6-xWO12-δ–Based Membranes. Chem. Mater. Res. 2012, 2, 27–81.
  107. Yoshimura, M. X-Ray Characterization and Thermal Properties of 3R2O3•2WO3 Compounds (R = La, Ce, Pr, and Nd). J. Am. Ceram. Soc. 1977, 60, 77–78. [CrossRef]
  108. Shlyakhtina, A.; Lyskov, N.; Baldin, E.; Stolbov, D.; Kolbanev, I.; Shatov, A.; Kasyanova, A.; Medvedev, D. Impact of Ln Cation on the Oxygen Ion Conductivity of Ln14W4O33 (Ln = Nd, Sm, Gd, Dy, Ho, Er, Tm, Yb) Tungstates. Ceram. Int. 2024, 50, 704–713. [CrossRef]
  109. McCarthy, G.J.; Fischer, R.D.; Jonson, G.G.; Gooden, C.E. Crystal Chemistry and Compound Formation in the Systems Rare Earth Sesquioxide-WO3, In Solid State Chemistry; Roth, R.S., Schneider Jr., S.J., Eds.; National Bureau of Standards, Department of Commerce: Washington, D.C., USA, 1972; pp. 397–410.
  110. Nilsson, M.; Grins, J.; Käll, P.-O.; Svensson, G. Synthesis, Structural Characterisation and Magnetic Properties of Gd14W4O33−xNy (0 ⩽ x ⩽ 17 ± 2, 0 ⩽ y ⩽ 9 ± 2), a New Fluorite-Related Oxynitride. J. Alloys Compd. 1996, 240, 60–69. [CrossRef]
  111. Evans, I.R.; Howard, J.A.K.; Evans, J.S.O. The Crystal Structure of α-La2Mo2O9 and the Structural Origin of the Oxide Ion Migration Pathway. Chem. Mater. 2005, 17, 4074–4077. [CrossRef]
  112. López-Vergara, A.; Vizcaíno-Anaya, L.; Porras-Vázquez, J.M.; Baldinozzi, G.; dos Santos-Gómez, L.; Canales-Vazquez, J.; Marrero-López, D.; Losilla, E.R. Unravelling Crystal Superstructures and Transformations in the La6–xMoO12−δ (0.6 ≤ x ≤ 3.0) Series: A System with Tailored Ionic/Electronic Conductivity. Chem. Mater. 2020, 32, 7052–7062. [CrossRef]
  113. Shlyakhtina, A.V.; Lyskov, N.V.; Kolbanev, I.V.; Shchegolikhin, A.N.; Karyagina, O.K.; Shcherbakova, L.G. Key Trends in the Proton Conductivity of Ln6−xMoO12−δ (Ln = La, Nd, Sm, Gd -Yb; x = 0, 0.5, 0.6, 0.7, 1) Rare-Earth Molybdates. Int. J. Hydrog. Energy 2021, 46, 16989–16998. [CrossRef]
  114. Baldin, E.D.; Lyskov, N.V.; Rassulov, V.A.; Shlyakhtina, A.V. Phase Formation, Polymorphism, Optical Properties and Conductivity of Compounds and Solid Solutions Based on Nd2WO6. J. Phys. Chem. 2024, accepted.
  115. Geng, X.; Hang, G.; Fernández-Carrión, A.J.; Ming, X.; Deng, S.; He, L.; Kuang, X.; Yang, X. Mixed Oxide Ion-Proton Conductivity and the Ionic Migration Mechanism in Isolated Tetrahedral LaVO4 by Acceptor Doping. Inorg. Chem. Front. 2024, 11, 5014–5025. [CrossRef]
  116. Kharitonova, E.P.; Orlova, E.I.; Gorshkov, N.V.; Goffman, V.G.; Voronkova, V.I. Bi2O3–Nd2O3–WO3 System: Phase Formation, Polymorphism, and Conductivity. Ceram. Int. 2021, 47, 31168–31179. [CrossRef]
  117. Kharitonova, E.P.; Orlova, E.I.; Gorshkov, N.V.; Goffman, V.G.; Chernyak, S.A.; Voronkova, V.I. Polymorphism and conductivity of Bi2O3-based fluorite-like compounds in Bi2O3–Nd2O3–MoO3 system. J. Alloys Compd. 2019, 787, 452–462. [CrossRef]
  118. Kharitonova, E.P.; Orlova, E.I.; Gorshkov, N.V.; Goffman, V.G.; Voronkova, V.I. Stabilized Bi2O3-Based Phases in the Bi2O3–Pr2O3–MoO3 System and Their Electrical Properties. Ceram. Int. 2018, 44, 12886–12895. [CrossRef]
  119. Orlova, E.I.; Kharitonova, E.P.; Gorshkov, N.V.; Goffman, V.G.; Voronkova, V.I. Phase Formation and Electrical Properties of Bi2O3-Based Compounds in the Bi2O3-La2O3-MoO3 System. Solid State Ionics 2017, 302, 158–164. [CrossRef]
  120. Miyazaki, S.; Li, Z.; Li, L.; Toyao, T.; Nakasaka, Y.; Nakajima, Y.; Shimizu, K.; Maeno, Z. Chemical Looping Dry Reforming of Methane over Ni-Modified WO3/ZrO2: Cooperative Work of Dispersed Tungstate Species and Ni over the ZrO2 Surface. Energy Fuels 2023, 37, 7945–7957. [CrossRef]
  121. Devillers, M.; De Smet, F.; Tirion; O. Bismuth and Mixed Bismuth-Lanthanide Carboxylates as Precursors for Pure and Ln-Promoted Bismuth Molybdate Catalysts. Thermochim. Acta 1995, 260, 165–185. [CrossRef]
  122. Maione, A. Ruiz, P.; Devillers, M. Rationalization of the Role Played by Bismuth and Lanthanides in Modified Ni–Co Molybdates as Catalysts for Partial and Total Oxidation of Propane. Catal. Today 2004, 91–92, 121–125. [CrossRef]
  123. De Smet, F.; Ruiz, P.; Delmon, B.; Devillers, M. Rationalization of the Catalytic Behavior of Lanthanide Oxides and Praseodymium Molybdates in Total and Selective Oxidation of Isobutene. J. Phys Chem. B 2000, 105, 12355–12363. [CrossRef]
  124. Kilbourn, B.T. The Role of the Lanthanides in Applied Catalysis. J. Less-Common Met. 1986, 126, 101–106. [CrossRef]
  125. Shi, X.; Wang, Z.; Takei, T.; Wang, X.; Zhu, Q.; Li, X.; Kim. B.-N.; Sun, X.; Li, J.-G. Selective Crystallization of Four Tungstates (La2W3O12, La2W2O9, La14W8O45, and La6W2O15) via Hydrothermal Reaction and Comparative Study of Eu3+ Luminescence. Inorg. Chem. 2018, 57, 6632–6640. [CrossRef]
  126. Shlyakhtina, A.V.; Avdeev, M.; Lyskov, N.V.; Abrantes, J.C.C.; Gomes, E.; Denisova, K.N.; Kolbanev, I.V.; Chernyak, S.A.; Volkova, O.S.; Vasiliev, A.N. Structure, Conductivity and Magnetism of Orthorhombic and Fluorite Polymorphs in MoO3–Ln2O3 (Ln = Gd, Dy, Ho) Systems. Dalton Trans. 2020, 49, 2833–2842. [CrossRef]
  127. Baldin, E.; Lyskov, N.; Vorobieva, G.; Kolbanev, I.; Karyagina, O.; Stolbov, D.; Voronkova, V.; Shlyakhtina, A. Synthesis of Hexagonal Nanophases in the La2O3–MO3 (M = Mo, W) Systems. Energies 2023, 16, 5637. [CrossRef]
  128. 103Yoshimura, M.; Rouanet, A. High Temperature Phase Relation in the System La2O3-WO3. Mater. Res. Bull. 1976, 11, 151–158. [CrossRef]
  129. Varga T.; Moats, J.L.; Ushakov, S.V.; Navrotsky, A. Thermochemistry of A2M3O12 Negative Thermal Expansion Materials. J. Mater. Res. 2007, 22, 2512–2521. [CrossRef]
  130. Kireeva, N.; Tsivadze, A.Yu. Oxide Ceramics of A2M3O12 Family with Negative and Close-to-Zero Thermal Expansion Coefficients: Machine Learning-Based Modeling of Functional Characteristics. J. Alloys Compd. 2024, 990, 174356. [CrossRef]
  131. Forster, P.M.; Yokochi, A.; Sleight, A.W. Enhanced Negative Thermal Expansion in Lu2W3O12. J. Solid State Chem. 1998, 140, 157–158.
  132. Khaliullin, Sh.M.; Khaliullina, A.Sh. Electrical Conductivity in Y2W3O12 System. In Proceedings of the Physical Chemistry and Electrochemistry of Molten and Solid Electrolytes, Yekaterinburg, Russia, 16–20 September 2013; pp. 263–264. Available online: http://elar.urfu.ru/handle/10995/135001 (accessed on 31 July 2024).
  133. Sumithra, S.; Umarji, A.M. Role of Crystal Structure on the Thermal Expansion of Ln2W3O12 (Ln = La, Nd, Dy, Y, Er and Yb). Solid State Sci. 2004, 6, 1313–1319. [CrossRef]
  134. Liu, Q.; Fan, C.; Wu, G.; Zhao, Y.; Sun, X.; Cheng, X.; Shen, J.; Hu, Y. In-situ Synthesis of Sc2W3O12/YSZ Ceramic Composites with Controllable Thermal Expansion. Ceram. Int. 2015, 41, 8267–8271. [CrossRef]
  135. Guseva, A.F.; Pestereva, N.N.; Vostrotina, E.L.; Otcheskikh, D D.; Lopatin, D.A. Ionic Conductivity of Solid Solutions and Composites Based on Sm2W3O12. Russ. J. Electrochem. 2020, 56, 447–451. [CrossRef]
  136. Neiman, A.Ya.; Pestereva, N.N.; Nechaev, D.O.; Koteneva, E.A.; Volkova, N.A.; Korchuganova, I.G.; Vanec, K.; Higgins, B.; Zhou, Y. The Nature and the Mechanism of Ion Transfer in Tungstates Me2+{WO4} (Ca, Sr, Ba) and Me3+2{WO4}3 (Al, Sc, In) According to the Data Acquired by the Tubandt Method. Russ. J. Electrochem. 2013, 49, 895–907. [CrossRef]
  137. Pestereva, N.; Guseva, А.; Vyatkin, I.; Lopatin, D. Electrotransport in Tungstates Ln2(WO4)3 (Ln = La, Sm, Eu, Gd). Solid State Ionics 2017, 301, 72–77. [CrossRef]
  138. Liu, Q.; Yang, J.; Cheng, X.; Liang, G.; Sun, X. Preparation and Characterization of Negative Thermal Expansion Sc2W3O12/Cu Core–Shell Composite. Ceram. Int. 2012, 38, 541–545. [CrossRef]
  139. Zhang, Z.; Tian, H.; Zhang, C.; Zhou, Y.; Wang, W.; Meng, X.; Liu, H. Flower-Like Sc2Mo3O12 Nanosheet Clusters: Synthesis, Thermal Expansion and Photocatalytic Properties. Ceram. Int. 2022, 48, 30135–30143. [CrossRef]
  140. Kuriakose, S.; H, H.; Jose, A.; John, M.; Varghese, T. Structural and Optical Characterization of Lanthanum Tungstate Nanoparticles Synthesized by Chemical Precipitation Route and Their Photocatalytic Activity. Opt. Mater. 2020, 99, 109571. [CrossRef]
  141. Zhou, Y.; Adams, S.; Rao, R.P.; Edwards, D.D.; Neiman, A.; Pestereva, N. Charge Transport by Polyatomic Anion Diffusion in Sc2(WO4)3. Chem. Mater. 2008, 20, 6335–6345. [CrossRef]
  142. Zhou, Y.; Rao, R.P.; Adams, S. Mechanism of Defect Formation and Polyanion Transport in Solid Scandium Tungstate Type Oxides. Monatsh. Chem. 2009, 140, 1017–1023. [CrossRef]
  143. Zhou, Y.; Rao, R.P.; Adams, S. Intrinsic Polyatomic Defects in Sc2(WO4)3. Solid State Ionics 2011, 192, 34–37. [CrossRef]
  144. Chellappa, A.S.; Viswanath, D.S. Partial Oxidation of Methane Using Ferric Molybdate Catalyst. Ind. Eng. Chem. Res. 1995, 34, 1933–1940. [CrossRef]
  145. Otsuka, K.; Wang, Y.; Yamanaka, I.; Morikawa, A.; Sinev, M.Yu. Partial Oxidation of Methane over Iron Molybdate Catalyst. Stud. Surf. Sci. Catal. 1994, 81, 503–508. [CrossRef]
  146. Arab, M.; Lopes-Moriyama, A.L.; dos Santos, T.R.; de Souza, C.P.; Gavarri, J.R.; Leroux, C. Strontium and Cerium Tungstate Materials SrWO4 and Ce2(WO4)3: Methane Oxidation and Mixed Conduction. Catal. Today 2013, 208, 35–41. [CrossRef]
  147. Oudghiri-Hassani, H. Synthesis, Characterization and Catalytic Performance of Iron Molybdate Fe2(MoO4)3 Nanoparticles. Catal. Commun. 2015, 60, 19–22. [CrossRef]
  148. Cheng, Y.Z.; Xue, L.; Sun, X.Y.; Hu, Z.B. Effects of Rare Earth Dopants on Crystal Structures of Bismuth Molybdate Catalysts. Adv. Mater. Res. 2011, 295–297, 1046–1049. [CrossRef]
  149. Snyder, T.P., Hill, C.G. The Mechanism for the Partial Oxidation of Propylene over Bismuth Molybdate Catalysts. Catal. Rev. 1989, 31, 43–95. [CrossRef]
  150. Machiels, C.J.; Chowdhry, U.; Harrison, W.T.A.; Sleight, A.W. Molybdate and Tungstate Catalysts for Methanol Oxidation. ACS Symp. Ser. 1985, 279, 103–119. [CrossRef]
  151. Porotnikova, N.; Khrustov, A.; Farlenkov, A.; Khodimchuk, A.; Partin, G.; Animitsa, I.; Kochetova, N.; Pavlov, D.; Ananyev, M. Promising La2Mo2O9–La2Mo3O12 Composite Oxygen-Ionic Electrolytes: Interphase Phenomena. ACS Appl. Mater. Interfaces 2022, 14, 6180–6193. [CrossRef]
  152. Shlyakhtina, A.V.; Baldin, E.D.; Vorobieva, G.A.; Kolbanev, I.V.; Stolbov, D.N.; Kasyanova, A.V.; Lyskov, N.V. Proton/Oxygen Ion Conductivity Ratio of Nd Containing La10W2O21/γ-La6W2O15 Tungstates. Int. J. Hydrog. Energy 2023, 48, 22671–22684. [CrossRef]
  153. Orlova, E.I.; Morkhova, Ye.A.; Egorova, A.V.; Kabanov, A.A.; Baldin, E.D.; Kharitonova, E.P.; Lyskov, N.V.; Yapaskurt, V.O.; Alekseeva, O.A.; Voronkova, V.I.; Korona, D.V. Extensive Research of Conductivity in the Fluorite-Like KLn4Mo3O15F (Ln = La, Pr, Nd) Rare Earth Molybdates: Theoretical and Experimental Data. Phys. Chem. Chem. Phys. 2024, 26, 7772–7782. [CrossRef]
  154. Bréard, Y.; Michel, C.; Hervieu, M.; Raveau, B. A Mixed Valent Iron Oxycarbonate Closely Related to the n = 3 Member of the RP Series Sr4Fe3−x(CO3)xO10−4xδ. J. Mater. Chem. 2000, 10, 1043−1045. [CrossRef]
  155. Tachibana, S.; Zhong, C.; Ide, K.; Yamasaki, H.; Tojigamori, T.; Miki, H.; Saito, T.; Kamiyama, T.; Shimoda, K.; Orikasa, Y. Fluorosulfide La2+xSr1−xF4+xS2 with a Triple-Fluorite Layer Enabling Interstitial Fluoride-Ion Conduction. Chem. Mater. 2023, 35, 4235−4242. [CrossRef]
  156. Ueno, N.; Yaguchi, H.; Fujii, K.; Yashima, M. High Conductivity and Diffusion Mechanism of Oxide Ions in Triple Fluorite-Like Layers of Oxyhalides. J. Am. Chem. Soc. 2024, 146, 11235−11244. [CrossRef]
  157. Araújo, A.J.M.; Macedo, D.A.; Graça, V.C.D.; Holz, L.I.V.; Fagg, D.P.; Loureiro, F.J.A. Progress in Misfit Ca-Cobaltite Electrodes for Solid Oxide Electrochemical Cells. In Handbook of Energy Materials; Gupta, R., Ed.; Springer: Singapore, 2022; pp. 1–34. [CrossRef]
  158. Pikalova, E.; Kolchugin, A.; Koroleva, M.; Vdovin, G.; Farlenkov, A.; Medvedev, D. Functionality of an Oxygen Ca3Co4O9+δ Electrode for Reversible Solid Oxide Electrochemical Cells Based on Proton-Conducting Electrolytes. J. Power Sources 2019, 438, 226996.
  159. Filonova, E.A.; Tokareva, E.S.; Pikalova, N.S.; Vylkov, A.I.; Bogdanovich, N.M.; Pikalova, E.Yu. Assessment of Prospective Cathodes Based on (1-x)Ca3Co4O9+δ-xBaCe0.5Zr0.3Y0.1Yb0.1O3-δ Composites for Protonic Ceramic Electrochemical Cells. J. Solid State Electrochem. 2020, 24, 1509–1521. [CrossRef]
  160. Thoréton, V.; Hu, Y.; Pirovano, C.; Capoen, E.; Nuns, N.; Mamede, A.S.; Dezanneau, G.; Yoo, C.Y.; Bouwmeester, H.J.M.; Vannier, R.N. Oxygen Transport Kinetics of the Misfit Layered Oxide Ca3Co4O9+δ. J. Mater. Chem. A 2014, 2, 19717–19725. [CrossRef]
  161. Zhang, Y.; Ohta, H. Recent Progress in Thermoelectric Layered Cobalt Oxide Thin Films. NPG Asia Mater. 2023, 15, 67. [CrossRef]
  162. Zhang, L.; Liu, Y.; Tan, T.T.; Liu, Y.; Zheng, J.; Yang, Y.; Hou, X.; Feng, L.; Suo, G.; Ye, X.; Li, S. Thermoelectric Performance Enhancement by Manipulation of Sr/Ti Doping in Two Sublayers of Ca3Co4O9. J. Adv. Ceram. 2020, 9, 769–781. [CrossRef]
  163. Nagasawa, K.; Daviero-Minaud, S.; Preux, N.; Rolle, A.; Roussel, P.; Nakatsugawa, H.; Mentré, O. Ca3Co4O9−δ: A Thermoelectric Material for SOFC Cathode. Chem. Mater. 2009, 21, 4738–4745. [CrossRef]
  164. Padmasree, K.P.; Lai, K.-Y.; Manthiram, A. Synthesis and Characterization of Ca3-xLaxCo4-yCuyO9+δ Cathodes for Intermediate Temperature Solid Oxide Fuel Cells. Ceram. Int. 2022, 48, 455–462. [CrossRef]
  165. Saqib, M.; Choi, I.-G.; Bae, H.; Park, K.; Shin, J.-S.; Kim, Y.-D.; Lee, J.-I.; Jo, M.; Kim, Y.-C.; Lee, K.-S.; Song, S.-J.; Wachsman, E.D.; Park, J.-Y. Transition from Perovskite to Misfit-Layered Structure Materials: A Highly Oxygen Deficient and Stable Oxygen Electrode Catalyst. Energy Environ. Sci. 2021, 14, 2472–2484. [CrossRef]
  166. Loureiro, F.J.A.; Silva, V.D.; Simões, T.A.; Cesário, M.R.; Grilo, J.P.F.; Fagg, D.P.; Macedo, D.A. Misfit-Layered Ca-Cobaltite–Based Cathodes for Intermediate-Temperature Solid Oxide Fuel Cell. In Intermediate Temperature Solid Oxide Fuel Cells. Electrolytes, Electrodes and Interconnects; Kaur, G., Ed.; Elsevier, 2020; pp. 347–377. [CrossRef]
  167. Yurchenko, M.V.; Antonova, E.P.; Tropin, E.S.; Suntsov, A.Yu. Adjusting Electrochemical Properties of PrBaCo2O6–δ as SOFC Cathode by Controllable Ca3Co4O9 Additions. Ceram. Int. 2023, 49, 21485–21491. [CrossRef]
  168. An, W.Y.; Kim, S.; Lee, W.; Choi, S.; Choi, S.R.; Yoo, S.; Han, J.W.; Li, O.L.; Park, J.-Y. Electrically and Morphologically Tailored Misfit-Layered Structure Gd0.3Ca2.7Co3.82Cu0.18O9 Nanofibers as Efficient Oxygen Catalysts for Zinc-Air Batteries. Appl. Catal. B Environ. Energy 2024, 358, 124354. [CrossRef]
  169. Yu, S.; He, S.; Chen, H.; Guo, L. Effect of Calcination Temperature on Oxidation State of Cobalt in Calcium Cobaltite and Relevant Performance as Intermediate-Temperature Solid Oxide Fuel Cell Cathodes. J. Power Sources 2015, 280, 581–587. [CrossRef]
  170. Kang, M.-G.; Cho, K.-H.; Kim, J.-S.; Nahm, S.; Yoon, S.-J.; Kang, C.-Y. Post-Calcination, a Novel Method to Synthesize Cobalt Oxide-Based Thermoelectric Materials. Acta Mater. 2014, 73, 251–258. [CrossRef]
  171. Madre, M.A.; Costa, F.M.; Ferreira, N.M.; Sotelo, A.; Torres, M.A.; Constantinescu, G.; Rasekh, Sh.; Diez, J.C. Preparation of High-Performance Ca3Co4O9 Thermoelectric Ceramics Produced by a New Two-Step Method. J. Eur. Ceram. Soc. 2013, 33, 1747–1754. [CrossRef]
  172. Zhang, Y.; Zhang, J.; Lu, Q. Synthesis of Highly Textured Ca3Co4O9 Ceramics by Spark Plasma Sintering. Ceram. Int. 2007, 33, 1305–1308. [CrossRef]
  173. dos Santos, A.M.; Thomazini, D.; Gelfuso, M.V. Cold Sintering and Thermoelectric Properties of Ca3Co4O9 Ceramics. Ceram. Int. 2020, 46, 14064–14070. [CrossRef]
  174. Constantinescu, G.; Rasekh, Sh.; Torres, M.A.; Bosque, P.; Diez, J.C.; Madre, M.A.; Sotelo, A. Effect of Na Doping on the Ca3Co4O9 Thermoelectric Performance. Ceram. Int. 2015, 41, 10897–10903. [CrossRef]
  175. Bae, H.; Kim, I.-H.; Park, H.-K.; Park, J.-Y.; Song, S.-J. Nonideal Defect Structure and High-Temperature Transport Properties of Misfit-Layered Cobalt Oxide. J. Solid State Chem. 2022, 313, 123299. [CrossRef]
  176. Fang, J.; Yang, H.; Liu, L.; Kang, Q.; Gou, Y. Research Progress on Doping Modification of Ca3Co4O9 Thermoelectric Materials: A Review. J. Mater. Sci. 2024, 59, 2228–2257. [CrossRef]
  177. Qi, X.L.; Fan, Y.Y.; Zeng, L.K.; Zhu, D.S. Effects of Mg Substitution on the Thermoelectric Properties of Ca3Co4O9-Based Materials. AMM. 2011, 84–85, 671–675. [CrossRef]
  178. Zhang, F.; Lu, Q.; Li, T.; Zhang, X.; Zhang, J.; Song, X. Preparation and Thermoelectric Transport Properties of Ba-, La- and Ag-Doped Ca3Co4O9 Oxide Materials. J. Rare Earths 2013, 31, 778–783. [CrossRef]
  179. Park, K.; Park, G.-M.; Park, J.-Y. Modification of Misfit-Layered Cathode Materials for Improving Electrocatalyic Activity in Reversible Protonic Ceramic Cells. ECS Trans. 2023, 111, 1007–1012. [CrossRef]
  180. Park, K.; Bae, H.; Kim, H.-K.; Choi, I.-G.; Jo, M.; Park, G.-M.; Asif, M.; Bhardwaj, A.; Lee, K.-S.; Kim, Y.-C.; Song, S.-J.; Wachsman, E.D.; Park, J.-Y. Understanding the Highly Electrocatalytic Active Mixed Triple Conducting NaxCa3–xCo4O9–δ Oxygen Electrode Materials. Adv. Energy Mater. 2023, 13, 2202999. [CrossRef]
  181. Klyndyuk, A.I.; Matsukevich, I.V. Synthesis and Properties of Ca2.8Ln0.2Co4O9+δ (Ln = La, Nd, Sm, Tb-Er) Solid Solutions. Inorg. Mater. 2012, 48, 1052–1057. [CrossRef]
  182. Ben Yahia, H.; Mauvy, F.; Grenier, J.C. Ca3−xLaxCo4O9+δ (x=0, 0.3): New Cobaltite Materials as Cathodes for Proton Conducting Solid Oxide Fuel Cell. J. Solid State Chem. 2010, 183, 527–531. [CrossRef]
  183. Wang, Y.; Sui, Y.; Cheng, J.; Wang, X.; Su, W.; Fan, H. Influence of Y3+ Doping on the High-Temperature Transport Mechanism and Thermoelectric Response of Misfit-Layered Ca3Co4O9. Appl. Phys. A 2010, 99, 451–458. [CrossRef]
  184. Urusova, A.; Bryuzgina, A.; Solomakhina, E.; Kolchugin, A.; Malyshkin, D.; Pikalova, E.; Filonova, E. Assessment of the Y-Doped Ca3Co4O9+δ as Cathode Material for Proton-Conducting Fuel Cells. Int. J. Hydrog. Energy 2023, 48, 22656–22670. [CrossRef]
  185. Cho, J.Y.; Kwon, O.J.; Chung, Y.K.; Kim, J.-S.; Kin, W.-S.; Song, K.J.; Park, C. Effect of Trivalent Bi Doping on the Seebeck Coefficient and Electrical Resistivity of Ca3Co4O9. J. Electron. Mater. 2015, 44, 3621–3626. [CrossRef]
  186. Boyle, C.; Carvillo, P.; Chen, Y.; Barbero, E.J.; Mcintyre, D.; Song, X. Grain Boundary Segregation and Thermoelectric Performance Enhancement of Bismuth Doped Calcium Cobaltite. J. Eur. Ceram. Soc. 2016, 36, 601–607. [CrossRef]
  187. Wang, Y.; Sui, Y.; Cheng, J.; Wang, X.; Su, W. Comparison of the High Temperature Thermoelectric Properties for Ag-Doped and Ag-Added Ca3Co4O9. J. Alloys Compd. 2009, 477, 817–821. [CrossRef]
  188. Lima, C.G.M.; Silva, R.M.; de M. Aquino, F.; Raveau, B.; Caignaert, V.; Cesário, M.R.; Macedo, D.A. Proteic Sol-Gel Synthesis of Copper Doped Misfit Ca-Cobaltites with Potential SOFC Application. Mater. Chem. Phys. 2017, 187, 177–182. [CrossRef]
  189. Yue, Y.; Yu, S.; Gu, Y.; Bi, L. A New Fe-Doped Ca3Co4O9 Cathode for Protonic Ceramic Fuel Cells. Ceram. Int. 2024, in press. [CrossRef]
  190. Zhang, D.; Mi, X.; Wang, Z.; Tang, G.; Wu, Q. Suppression of the Spin Entropy in Layered Cobalt Oxide Ca3Co4O9+δ by Cu Doping. Ceram. Int. 2014, 40, 12313–12318. [CrossRef]
  191. Wang, Y.; Sui, Y.; Wang, X.; Su, W.; Liu, X. Enhanced High Temperature Thermoelectric Characteristics of Transition Metals Doped Ca3Co4O9+δ by Cold High-Pressure Fabrication. J. Appl. Phys. 2010, 107, 033708. [CrossRef]
  192. Yao, Q.; Wang, D.L.; Chen, L.D.; Shi, X.; Zhou, M. Effects of Partial Substitution of Transition Metals for Cobalt on the High-Temperature Thermoelectric Properties of Ca3Co4O9+δ. J. Appl. Phys. 2005, 97, 103905. [CrossRef]
  193. Wang, S.-F.; Hsu, Y.-F.; Chang, J.-H.; Cheng, S.; Lu, H.-C. Characteristics of Cu and Mo-Doped Ca3Co4O9−δ Cathode Materials for Use in Solid Oxide Fuel Cells. Ceram. Int. 2016, 42, 11239–11247. [CrossRef]
  194. Prasoetsopha, N.; Pinitsoontorn, S.; Kamwanna, T.; Kurosaki, T.; Ohishi, Y.; Muta, H.; Yamanaka, S. Thermoelectric Properties of Ca3Co4-xGaxO9+δ Prepared by Thermal Hydro-decomposition. J. Electron. Mater. 2014, 43, 2064–2071. [CrossRef]
  195. Tian, R.; Donelson, R.; Ling, C.D.; Blanchard, P.E.R.; Zhang, T.; Chu, D.; Tan, T.T.; Li, S. Ga Substitution and Oxygen Diffusion Kinetics in Ca3Co4O9+δ-Based Thermoelectric Oxides. J. Phys. Chem. C 2013 117, 13382–13387. [CrossRef]
  196. Klyndyuk, A.I.; Matsukevich, I.V. Synthesis, Structure, and Properties of Ca3Co3.85M0.15O9+δ (M = Ti–Zn, Mo, W, Pb, Bi) Layered Thermoelectrics. Inorg. Mater. 2015, 51, 944–950. [CrossRef]
  197. Xu, L.; Li, F.; Wang, Y. High-Temperature Transport and Thermoelectric Properties of Ca3Co4−xTixO9. J. Alloys Compd. 2010, 501, 115–119. [CrossRef]
  198. Xu, W.; Butt, S.; Zhu, Y.; Zhou, J.; Liu, Y.; Yu, M.; Marcelli, A.; Lan, J.; Lin, Y.-H.; Nan, C.-W. Nanoscale Heterogeneity in Thermoelectrics: The Occurrence of Phase Separation in Fe-doped Ca3Co4O9. Phys. Chem. Chem. Phys. 2016, 18, 14580–14587. [CrossRef]
  199. Hira, U.; Shahbaz Ali, S.; Latif, S.; Pryds, N.; Sher, F. Improved High-Temperature Thermoelectric Properties of Dual-Doped Ca3Co4O9. ACS Omega 2022, 7, 6579–6590. [CrossRef]
  200. Ou, Y.; Peng, J.; Li, F.; Yu, Z.X.; Ma, F.Y.; Xie, S.H.; Li, J.-F.; Li, J.Y. The Effects of Dual Doping on the Thermoelectric Properties of Ca3−xMxCo4−yCuyO9 (M = Na, La). J. Alloys Compd. 2012, 526, 139–144. [CrossRef]
  201. Wu, N. Nong, N.V.; Pryds, N.; Linderoth, S. Effects of Yttrium and Iron Co-Doping on the High Temperature Thermoelectric Properties of Ca3Co4O9+δ. J. Alloys Compd. 2015, 638, 127–132. [CrossRef]
  202. Zou, J.; Park, J.; Yoon, H.; Kim, T.; Chung, J. Preparation and Evaluation of Ca3−xBixCo4O9−δ (0< x ≤ 0.5) as Novel Cathodes for Intermediate Temperature-Solid Oxide Fuel Cells. Int. J. Hydrog Energy 2012, 37, 8592–8602. [CrossRef]
  203. Rolle, A.; Thoréton, V.; Rozier, P.; Capoen, E.; Mentré, O.; Boukamp, B.; Daviero-Minaud, S. Evidence of the Current Collector Effect: Study of the SOFC Cathode Material Ca3Co4O9+δ. Fuel Cells 2012, 12, 288–301. [CrossRef]
  204. Santos, J.R.D.; Loureiro, F.J.A.; Grilo, J.P.F.; Silva, V.D.; Simões, T.A.; Macedo, D.A. Understanding the Cathodic Polarisation Behaviour of the Misfit [Ca2CoO3−δ]q[CoO2] (C349) as Oxygen Electrode for IT-SOFC. Electrochim. Acta 2018, 285, 214–220. [CrossRef]
  205. Araújo, A.J.M.; Loureiro, F.J.A.; Holz, L.I.V.; Graça, V.C.D.; Grilo, J.P.F.; Macedo, D.A.; Paskocimas, C.A.; Fagg, D.P. Boosted Electrochemical Performance of Ca-Cobaltite-Based Composite Electrodes for Reversible Solid Oxide Cells. Int. J. Energy. Res. 2022, 46, 22070–22077. [CrossRef]
  206. Loureiro, F.J.A.; Araújo, A.J.M.; Paskocimas, C.A.; Macedo, D.A.; Fagg, D.P. Polarisation Mechanism of the Misfit Ca-Cobaltite Electrode for Reversible Solid Oxide Cells. Electrochim. Acta 2021, 373, 137928. [CrossRef]
  207. Hu, Y.; Thoréton, V.; Pirovano, C.; Capoen, E.; Bogicevic, C.; Nuns, N.; Mamede, A.-S.; Dezanneau G.; Vannier, R.N. Oxide Diffusion in Innovative SOFC Cathode Materials. Faraday Discuss. 2014, 176, 31–47. [CrossRef]
  208. Zhu, X.; A, L.; Zhu, C.; Wang, Y.; Jin, J. Performance Evaluation of Ca3Co4O9-δ Cathode on Sm0.075Nd0.075Ce0.85O2-δ Electrolyte for Solid Oxide Fuel Cells. J. Alloys Compd. 2017, 694, 877–883. [CrossRef]
  209. Araújo, A.J.M.; Loureiro, F.J.A.; Holz, L.I.V.; Grilo, J.P.F.; Macedo, D.A.; Paskocimas, C.A.; Fagg, D.P. Composite of Calcium Cobaltite with Praseodymium-Doped Ceria: A Promising New Oxygen Electrode for Solid Oxide Cells. Int. J. Hydrog. Energy 2021, 46, 28258–28269. [CrossRef]
  210. Fulgêncio, E.B.G.A.; Loureiro, F.J.A.; Melo, K.P.V.; Silva, R.M.; Fagg, D.P.; Campos, L.F.A.; Macedo, D.A. Boosting the Oxygen Reduction Reaction of the Misfit [Ca2CoO3-δ]q[CoO2] (C349) by the Addition of Praseodymium Oxide. J. Alloy Compd. 2019, 788, 148–154. [CrossRef]
  211. Guo, P.; Huang, X.; Zh, X.; Lǚ, Z.; Zhou, Y.; Li, L.; Li, Z.; Wei, B.; Zhang, Y.; Su, W. A New Composite Material Ca3Co4O9+δ+La0.7Sr0.3CoO3 Developed for Intermediate-Temperature SOFC Cathode. Fuel Cells 2013, 13, 666–672. [CrossRef]
  212. Li, Y.; Zhang, X.; Tian, Y.; Wu, X.; Wang, L.; Zhu, J. Study of Ca3Co4O9+δ Oxygen Electrode with La0.6Sr0.4FeO3-δ Interlayer in YSZ-Based Reversible Solid Oxide Cells. Int. J. Hydrog. Energy 2024, 50, 441–449. [CrossRef]
  213. Dang, C.; Li, Y.; Yusuf, S.M.; Cao, Y.; Wang, H.; Yu, H.; Peng, F.; Li, F. Calcium Cobaltate: A Phase-Change Catalyst for Stable Hydrogen Production from Bio-glycerol. Energy Environ. Sci. 2018, 11, 660–668. [CrossRef]
  214. Sun, J.; Yang, M.; Li, G.; Yang, T.; Liao, F.; Wang, Y.; Xiong, M.; Lin, J. New Barium Cobaltite Series Ban+1ConO3n+3(Co8O8): Intergrowth Structure Containing Perovskite and CdI2-Type Layers. Inorg. Chem. 2006, 45, 9151–9153. [CrossRef]
  215. Jo, M.; Bae, H.; Park, K.; Hamayun, M.A.; Park, G.-M.; Kim, J.H.; Lee, K.T.; Lee, K.-S.; Song, S.-J.; Park, J.-Y. Layered Barium Cobaltite Structure Materials Containing Perovskite and CdI2-Based Layers for Reversible Solid Oxide Cells with Exceptionally High Performance. Chem. Eng. J. 2023, 451, 138954. [CrossRef]
  216. Rolle, A.; Preux, N.; Ehora, G.; Mentré, O.; Daviero-Minaud, S. Potentiality of Ba2Co9O14 as Cathode Material for IT-SOFC on Various Electrolytes. Solid State Ionics 2011, 184, 31–34. [CrossRef]
  217. Araújo, A.J.M.; Loureiro, F.J.A.; Raimundo, R.A.; Macedo, D.A.; Paskocimas, C.A.; Fagg, D.P. The Effects of Polarisation on the Performance of the Ba2Co9O14–Ce0.8Gd0.2O2-δ Composite Electrode for Fuel Cells and Electrolysers. Int. J. Hydrog. Energy 2022, 47, 11270–11278. [CrossRef]
  218. Araújo, A.J.M.; Graça, V.C.D.; Raimundo, R.A.; Filho, A.C.L.; Macedo, D.A.; Loureiro, F.J.A. A New Layered Barium Cobaltite Electrode for Protonic Ceramic Cells. J. Mater. Chem. A 2024, 12, 840–853. [CrossRef]
  219. Huang, D.; Liu, T.; Xu, A.; Zhou, J.; Wang, Y.; Hu, X. A Novel Layered Cobalt Oxide Ba2Co9O14 as an Efficient and Durable Bifunctional Oxygen Electrocatalyst for Rechargeable Zinc-Air Batteries. Electrochim. Acta 2024, 494, 144450. [CrossRef]
  220. Razavi, F.S.; Hajizadeh-Oghaz, M.; Amiri, O.; Morassaei, M.S.; Salavati-Niasari, M. Barium Cobaltite Nanoparticles: Sol-Gel Synthesis and Characterization and Their Electrochemical Hydrogen Storage Properties. Int. J. Hydrog. Energy 2021, 46, 886–895. [CrossRef]
  221. Bermanec, V.; Holtstam, D.; Sturman, D.; Criddle, A.J.; Back, M.E.; Šćavničar, S. Nežilovite, a New Member of the Magnetoplumbite Group, and the Crystal Chemistry of Magnetoplumbite and Hibonite. Canad. Mineral. 1996, 34, 1287–1297.
  222. Vinnik, D.A.; Trofimov, E.A.; Zhivulin, V.E.; Zaitseva, O.V.; Gudkova, S.A.; Starikov, A.Yu.; Zherebtsov, D.A.; Kirsanova, A.A.; Häßner, M.; Niewa, R. High-Entropy Oxide Phases with Magnetoplumbite Structure. Ceram. Int. 2019, 45, 12942–12948. [CrossRef]
  223. Wagner, T.R. Preparation and Crystal Structure Analysis of Magnetoplumbite-Type BaGa12O19. J. Solid State Chem. 1998, 136, 120–124. [CrossRef]
  224. Chukanov, N.V.; Vorobei, S.S.; Ermolaeva, V.N.; Varlamov, D.A.; Plechov, P.Y.; Jančev, S.; Bovkun, A.V. New Data on Chemical Composition and Vibrational Spectra of Magnetoplumbite-Group Minerals. Geol. Ore Depos. 2019, 61, 637–646. [CrossRef]
  225. Haoran, L.; Chang-An, W.; Chenguang, Z.; Shuyan, T. Thermo-Physical Properties of Rare-Earth Hexaaluminates LnMgAl11O19 (Ln: La, Pr, Nd, Sm, Eu and Gd) Magnetoplumbite for Advanced Thermal Barrier Coatings. J. Eur. Ceram. Soc. 2015, 35, 1297–1306. [CrossRef]
  226. Holtstam, D.; Hålenius, U. Nomenclature of the Magnetoplumbite Group. Mineral. Mag. 2020, 84, 376–380. [CrossRef]
  227. Jean, M.; Nachbaur, V.; Bran, J.; Le Breton, J.-M. Synthesis and Characterization of SrFe12O19 Powder Obtained by Hydrothermal Process. J. Alloys Compd. 2010, 496, 306–312. [CrossRef]
  228. Lu, Y.; Noor, A.; Ahmed, J.; Alwadie, N.; Akhtar, M.N.; Abid, S.; Yousaf, M.; Mahmoud, M.; Aslam, M. Synergistic Effects and Electrocatalytic Insight of Single-Phase Hexagonal Structure as Low-Temperature Solid Oxide Fuel Cell Cathode. J. Rare Earths 2024, in press. [CrossRef]
  229. Yamada, I.; Toda, F.; Kawaguchi, S.; Yagi, S. Multiple Factors on Catalytic Activity for Oxygen Evolution Reaction in Magnetoplumbite Fe–Co Oxide BaFe12–xCoxO19. ACS Appl. Energy Mater. 2022, 5, 5995–6002. [CrossRef]
  230. He, J.; Yang, Q.; Song, Z.; Chang, W.; Huang, C.; Zhu, Y.; Ma, X.; Wang, X. Improving the Carbon Resistance of Iron-Based Oxygen Carrier for Hydrogen Production via Chemical Looping Steam Methane Reforming: A Review. Fuel 2023, 351, 128864. [CrossRef]
  231. Tian, M.; Wang, X.D.; Zhang, T. Hexaaluminates: A Review of the Structure, Synthesis and Catalytic Performance. Catal. Sci. Technol. 2016, 6, 1984–2004. [CrossRef]
  232. Machida, M.; Sato, A.; Kijima, T.; Inoue, H.; Eguchi, K.; Arai, H. Catalytic Properties and Surface Modification of Hexaaluminate Microcrystals for Combustion Catalyst. Catal. Today 1995, 26, 239–245. [CrossRef]
  233. 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. [CrossRef]
  234. Závêta, K. Anisotropy of the Electrical Conductivity of Ferrites with the Magnetoplumbite Structure. Phys. Stat. Sol. (b) 1963, 3, 2111–2118. [CrossRef]
  235. Pullar, R.C. Hexagonal Ferrites: A Review of the Synthesis, Properties and Applications of Hexaferrite Ceramics. Prog. Mater. Sci. 2012, 57, 1191–1334. [CrossRef]
  236. Pal, M.; Brahma, P.; Chakraborty, B.R.; Chakravorty, D. DC Conductivity in Barium Hexaferrites Doped with Bismuth Oxide. Jpn. J. Appl. Phys. 1997, 36, 2163–2166. [CrossRef]
  237. Dudley, G.J.; Steele, B.C.H.; Howe, A.T. Studies of Potassium Ferrite K1+xFe11O17. I. Electronic Conductivity and Defect Structure. J. Solid State Chem. 1976, 18, 141–147. [CrossRef]
  238. Alamelu Mangai, K.; Tamizh Selvi, K.; Priya, M.; Rathnakumari, M.; Sureshkumar, P.; Sagadevan, S. Structural and Optical Absorption Studies of Cobalt Substituted Strontium Ferrites, SrCoxFe12−xO19 (x = 0.1, 0.2 and 0.3). J. Mater. Sci.: Mater. Electron. 2016, 28, 1238–1246. [CrossRef]
  239. Zhukovsky, V.M.; Bushkova, O.V.; Zainullina, V.M.; Dontsov, G.I.; Volosentseva, L.I.; Zhukovskaya, A.S. Diffusion Transport in Hexagonal Ferrites with Magnetoplumbite Structure. Solid State Ionics 1999, 119, 15–17. [CrossRef]
  240. Zainullina, V.M.; Zhukov, V.P.; Zhukovskii, V.M. Quantum-Chemical Calculation of the Electronic Structure and Ionic Conductivity of Lead Hexaferrite with a Magnetoplumbite Structure. J. Struct. Chem. 2001, 42, 705–710.
  241. Chen, X.; Zhang, Y.; Zhong, X.; Xu, Z.; Zhang, J.; Cheng, Y.; Zhao, Y.; Liu, Y.; Fan, X.; Wang, Y.; Ma, H.; Cao, X. Thermal Cycling Behaviors of the Plasma Sprayed Thermal Barrier Coatings of Hexaluminates with Magnetoplumbite Structure. J. Eur. Ceram. Soc. 2010, 30, 1649–1657. [CrossRef]
  242. Luo, X.; Huang, S.; Huang, R.; Hong, J.; Yuan, S.; Shu, Z.; Zhao, L.; Lu, C.; Jin, H. Rare-Earth High-Entropy Magnetoplumbite Structure Hexaluminates (La1/5Nd1/5Sm1/5Eu1/5Gd1/5)MAl11O19 (M = Mg, Zn) for Thermal Barrier Coating Applications with Enhanced Mechanical and Thermal Properties. Ceram. Int. 2024, 50, 21281–21288. [CrossRef]
  243. Butashin, A.V.; Chernyshev, A.P.; Muslimov, A.E.; Asvarov, A.Sh.; Kanevsky, V.M. Structural and Thermodynamic Aspects of the Evolution of Supersmooth Surface of Oxide Crystals. Appl. Surf. Sci. 2021, 553, 149541. [CrossRef]
  244. Gardner, T.H.; Spivey, J.J.; Campos, A.; Hissam, J.C.; Kugler, E.L.; Roy, A.D. Catalytic Partial Oxidation of CH4 over Ni-Substituted Barium Hexaaluminate Catalysts. Catal. Today 2010, 157, 166–169. [CrossRef]
  245. 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. AIChE J. 2018, 64, 550–563. [CrossRef]
  246. Wang, J.; Liu, Y.; Cheng, T.; Li, W.; Bi, L.; Zhen, K. Methane Reforming with Carbon Dioxide to Synthesis Gas over Co-Doped Ni-Based Magnetoplumbite Catalysts. Appl. Catal. A.: Gen. 2003, 250, 13–23. [CrossRef]
  247. Wang, J.; Meng, D.; Wu, X.; Hong, J., An, D.; Zhen, K. Catalytic Properties of Mg-Modified Ni-Based Hexaaluminate Catalysts for CO2 Reforming of Methane to Synthesis Gas. React. Kinet. Catal. Lett. 2009, 96, 65–73. [CrossRef]
  248. Karaismailoglu Elibol, M. Determination of Characteristic Properties of Co3O4 Loaded LaFexAl12−xO19 Hexaaluminates. Mater. Test. 2024, 66, 179–185. [CrossRef]
  249. Mill, B.V.; Pisarevsky, Y.V. Langasite-Type Materials: From Discovery to Present State. In Proceedings of the 2000 IEEE/EIA International Frequency Control Symposium and Exhibition (Cat. No.00CH37052), Kansas City, MO, USA, 09 June 2000; pp. 133–144. [CrossRef]
  250. Chai, B.H.T.; Bustamante, A.N.P.; Chou, M.C. A New Class of Ordered Langasite Structure Compounds. In Proceedings of the 2000 IEEE/EIA International Frequency Control Symposium and Exhibition (Cat. No.00CH37052), Kansas City, MO, USA, 09 June 2000; pp. 163–168. [CrossRef]
  251. Kuzmicheva, G.M.; Domoroschina, E.N.; Rybakov, V.B.; Dubovsky, A.B.; Tyunina, E.A. A Family of Langasite: Growth and Structure. J. Cryst. Growth 2005, 275, e715–e719. [CrossRef]
  252. Zhang, S.; Zheng, Y.; Kong, H.; Xin, J.; Frantz, E.; Shrout, T.R. Characterization of High Temperature Piezoelectric Crystals with an Ordered Langasite Structure. J. Appl. Phys. 2009, 105, 114107. [CrossRef]
  253. Diaz-Lopez, M.; Shin, J.F.; Li, M.; Dyer, M.S.; Pitcher, M.J.; Claridge, J.B.; Blanc, F.; Rosseinsky, M.J. Interstitial Oxide Ion Conductivity in the Langasite Structure: Carrier Trapping by Formation of (Ga,Ge)2O8 Units in La3Ga5−xGe1+xO14+x/2 (0 < x ≤ 1.5). Chem. Mater. 2019, 31, 5742–5758. [CrossRef]
  254. Anfimov, I.M.; Buzanov, O.A.; Kozlova, A.P.; Kozlova, N.S.; Zabelina, E.V. Impedance Spectroscopy Study of Lanthanum-Gallium Tantalate Single Crystals Grown under Different Conditions. Mod. Electron. Mater. 2019, 5, 41–49. [CrossRef]
  255. Luo, Z.; Li, X.; Wang, X.; Deng, S.; He, L.; Lin, K.; Li, Q.; Xing, X.; Kuang, X. The Langasite Family for the Development of Oxygen-Vacancy-Mediated Oxide Ion Conductors. Chem. Mater. 2024, 36, 2835–2845. [CrossRef]
  256. Corti, L.; Hung, I.; Venkatesh, A.; Gan, Z.; Claridge, J.B.; Rosseinsky, M.J.; Blanc, F. Cation Distribution and Anion Transport in the La3Ga5–xGe1+xO14+0.5x Langasite Structure. J. Am. Chem. Soc. 2024, 146, 14022–14035. [CrossRef]
  257. Seh, H.; Tuller, H.L. Defects and Transport in Langasite I: Acceptor-Doped (La3Ga5SiO14). J. Electroceram. 2006, 16, 115–125. [CrossRef]
  258. Bjørheim, T.S.; Haugsrud, R.; Norby, T. Protons in Acceptor Doped Langasite, La3Ga5SiO14. Solid State Ionics 2014, 264, 76–84. [CrossRef]
  259. Bjørheim, T.S.; Haugsrud, R. Proton Transport Properties of the RE3Ga5MO14 (RE = La, Nd and M = Si, Ti, Sn) Langasite Family of Oxides. Solid State Ionics 2015, 275, 29–34. [CrossRef]
  260. Bjørheim, T.S.; Shanmugappirabu, V.; Haugsrud, R.; Norby, T.E. Protons in Piezoelectric Langatate; La3Ga5.5Ta0.5O14. Solid State Ionics 2015, 278, 275–280. [CrossRef]
  261. Huminicki, D.M.C.; Hawthorne, F.C. Refinement of the Crystal Structure of Swedenborgite. Canad. Mineral. 2001, 39, 153–158.
  262. Huang, S.; Huang, Z.; Cao, P.; Zujovic, Z.; Price, J.R.; Avdeev, M.; Que, M.; Suzuki, F.; Kido, T.; Ouyang, X.; Kaji, H.; Fang, M.; Liu, Y.-G.; Gao, W.; Söhnel, T. “114”-Type Nitrides LnAl(Si4−xAlx)N7Oδ with Unusual [AlN6] Octahedral Coordination. Angew. Chem. Int. Ed. 2017, 56, 3886–3891. [CrossRef]
  263. Qureshi, N.; Ouladdiaf, B.; Senyshyn, A.; Caignaert, V.; Valldor, M. Non-Collinear Magnetic Structures in the Magnetoelectric Swedenborgite CaBaFe4O7 Derived by Powder and Single-Crystal Neutron Diffraction. SciPost Phys. Core 2022, 5, 007. [CrossRef]
  264. Ndubuisi, A.; Abouali, S.; Singh K.; Thangadurai, V. Recent Advances, Practical Challenges, and Perspectives of Intermediate Temperature Solid Oxide Fuel Cell Cathodes. J. Mater. Chem. A 2022, 10, 2196–2227. [CrossRef]
  265. Shin, J.-S.; Park, H.; Park, K.; Saqib, M.; Jo, M.; Kim, J.H.; Lim, H.-T.; Kim, M.; Kim, J.; Park, J.-Y. Activity of Layered Swedenborgite Structured Y0.8Er0.2BaCo3.2Ga0.8O7+δ for Oxygen Electrode Reactions in at Intermediate Temperature Reversible Ceramic Cells. J. Mater. Chem. A 2021, 9, 607–621. [CrossRef]
  266. Turkin, D.I.; Yurchenko, M.V.; Tolstov, K.S.; Shalamova, A.M.; Suntsov, A.Yu.; Kozhevnikov, V.L. Oxygen Exchange and Phase Stability of Y0.8Ca0.2BaCo4-xMxO7+δ (M = Fe, Ga, Al). J. Solid State Chem. 2023, 326, 124194. [CrossRef]
  267. Parkkima, O.; Karppinen, M. The YBaCo4O7+δ-Based Functional Oxide Material Family: A Review. Eur. J. Inorg. Chem. 2014, 2014, 4056–4067. [CrossRef]
  268. Kim, J.-H.; Manthiram, A. Low Thermal Expansion RBa(Co,M)4O7 Cathode Materials Based on Tetrahedral-Site Cobalt Ions for Solid Oxide Fuel Cells. Chem. Mater. 2010, 22, 822–831. [CrossRef]
  269. Kim, Y.N.; Kim, J.H.; Huq, A.; Paranthaman M.P.; Manthiram, A. (Y0.5In0.5)Ba(Co,Zn)4O7 Cathodes with Superior High-Temperature Phase Stability for Solid Oxide Fuel Cells. J. Power Sources 2012, 214, 7–14. [CrossRef]
  270. Lai, K.-Y.; Manthiram, A. Phase Stability, Oxygen-Storage Capability, and Electrocatalytic Activity in Solid Oxide Fuel Cells of (Y, In, Ca)BaCo4–yGayO7+δ. Chem. Mater. 2016, 28, 9077–9087. [CrossRef]
  271. Tang, C.; Yao, Y.; Wang, N.; Zhang, X.; Zheng, F.; Du, L.; Luo, D.; Aoki, Y.; Ye, S. Green Hydrogen Production by Intermediate-Temperature Protonic Solid Oxide Electrolysis Cells: Advances, Challenges, and Perspectives. InfoMat. 2024, 6, e12515. [CrossRef]
  272. Zhou, Q.; Zhang, X.; Wang, Y.; Ma, Y.; Yang, H. A Thermal-Expansion Offset to Cobalt-Based Cathode Materials for Solid Oxide Fuel Cells. Next Energy 2024, 5, 100168. [CrossRef]
  273. Zhou, Q.; Wang, Y.; Yang, X.; Bu, Fa.; Yang, F.; Wang, M.; Li, Y. Basic Properties of Low Thermal Expansion Coefficient (Y0.5Ca0.5)1−xInxBaCo3ZnO7+δ (x = 0, 0.1, 0.2, 0.3) Solid Solutions for Solid Oxide Fuel Cell Cathode Materials. Mater. Res. Bull. 2022, 156, 112001. [CrossRef]
  274. Ye, F.; Zhou, Q.; Xu, K.; Zhang, Z.; Han, X.; Yang, L.; Xu, J.; Xu, H.; Wu, K.; Guan, Y. Phase Stability and Electrochemical Performance of Y0.5Ca0.5−xInxBaCo3.2Ga0.8O7+δ (x = 0 and 0.1) as Cathodes for Intermediate Temperature Solid Oxide Fuel Cells. J. Alloys Compd. 2016, 680, 163–168. [CrossRef]
  275. Tsipis, E.V.; Khalyavin, D.D.; Shiryaev, S.V.; Redkina, K.S.; Núñez, P. Electrical and Magnetic Properties of YBaCo4O7+δ. Mater. Chem. Phys. 2005, 92, 33–38. [CrossRef]
  276. Tsipis, E.V.; Kharton, V.V.; Frade, J.R. Transport Properties and Electrochemical Activity of YBa(Co,Fe)4O7 Cathodes. Solid State Ionics 2006, 177, 1823–1826. [CrossRef]
  277. Avdeev, M.; Kharton, V.V.; Tsipis, E.V. Geometric Parameterization of the YBaCo4O7 Structure Type: Implications for Stability of the Hexagonal Form and Oxygen Uptake. J. Solid State Chem. 2010, 183, 2506–2509. [CrossRef]
Figure 1. Main types of conventional and state-of-the-art materials for SOFC/SOE, permselective membranes and catalysts for fuel transformation.
Figure 1. Main types of conventional and state-of-the-art materials for SOFC/SOE, permselective membranes and catalysts for fuel transformation.
Preprints 115443 g001
Figure 2. (a) Arrhenius plots for the oxygen ionic conductivity in Ln2MoO6 (Ln = Sm, Gd, Dy); (b) Results of the Density-Functional Theory (DFT) – Nudged Elastic Band (NEB) approach calculations of the low-energy oxygen diffusion paths for Dy2MoO6 [94]. Reprinted from Solid State Ionics, Vol. 400, Yelizaveta A. Morkhova et al., Comprehensive study of conductivity in the series of monoclinic oxymolybdates: Ln2MoO6 (Ln = Sm, Gd, Dy), Pages No. 116337, Copyright (2023), with permission from Elsevier.
Figure 2. (a) Arrhenius plots for the oxygen ionic conductivity in Ln2MoO6 (Ln = Sm, Gd, Dy); (b) Results of the Density-Functional Theory (DFT) – Nudged Elastic Band (NEB) approach calculations of the low-energy oxygen diffusion paths for Dy2MoO6 [94]. Reprinted from Solid State Ionics, Vol. 400, Yelizaveta A. Morkhova et al., Comprehensive study of conductivity in the series of monoclinic oxymolybdates: Ln2MoO6 (Ln = Sm, Gd, Dy), Pages No. 116337, Copyright (2023), with permission from Elsevier.
Preprints 115443 g002
Figure 3. Arrhenius plots for total conductivity in dry and wet air [105]: (1) hexagonal monocrystalline La2W1+xO6+3x (x ~ 0.22) [105]; (2) hexagonal ceramics La2W1+xO6+3x (x ~ 0.11) [105]; (3) hexagonal ceramics La2W1+xO6+3x (x ~ 0.11) according to the data [99], acquired in N2; (4) hexagonal ceramics La2W1+xO6+3x (x ~ 0.25) in air [103]; (5) rhombic ceramics β-La2WO6 [105]. Reprinted from A. V. Shlyakhtina et al, Specific Features of Phase Formation and Properties of Compounds La2W1+xO6+3x (x ~ 0; 0.11–0.22), Russian Journal of Electrochemistry, volume 59, pages 60 – 69, Copyright (2023), with permission from Springer Nature.
Figure 3. Arrhenius plots for total conductivity in dry and wet air [105]: (1) hexagonal monocrystalline La2W1+xO6+3x (x ~ 0.22) [105]; (2) hexagonal ceramics La2W1+xO6+3x (x ~ 0.11) [105]; (3) hexagonal ceramics La2W1+xO6+3x (x ~ 0.11) according to the data [99], acquired in N2; (4) hexagonal ceramics La2W1+xO6+3x (x ~ 0.25) in air [103]; (5) rhombic ceramics β-La2WO6 [105]. Reprinted from A. V. Shlyakhtina et al, Specific Features of Phase Formation and Properties of Compounds La2W1+xO6+3x (x ~ 0; 0.11–0.22), Russian Journal of Electrochemistry, volume 59, pages 60 – 69, Copyright (2023), with permission from Springer Nature.
Preprints 115443 g003
Figure 4. The total conductivity of the tungstate Ln14W4O33 (Ln = Nd–Yb) as a function of the partial pressure of oxygen at 900 °С [108]. Reprinted from Ceramics International, Vol. 50, Anna Shlyakhtina et al, Impact of Ln cation on the oxygen ion conductivity of Ln14W4O33 (Ln = Nd, Sm, Gd, Dy, Ho, Er, Tm, Yb) tungstates, Pages No. 704-713, Copyright (2024), with permission from Elsevier.
Figure 4. The total conductivity of the tungstate Ln14W4O33 (Ln = Nd–Yb) as a function of the partial pressure of oxygen at 900 °С [108]. Reprinted from Ceramics International, Vol. 50, Anna Shlyakhtina et al, Impact of Ln cation on the oxygen ion conductivity of Ln14W4O33 (Ln = Nd, Sm, Gd, Dy, Ho, Er, Tm, Yb) tungstates, Pages No. 704-713, Copyright (2024), with permission from Elsevier.
Preprints 115443 g004
Figure 7. Temperature dependencies of ionic conductivity for M2W3O12, M = Y [132] (a), Sc [133] (b), La [137] (c), Sm [137] (d), Eu [137] (e), Gd [137] (f).
Figure 7. Temperature dependencies of ionic conductivity for M2W3O12, M = Y [132] (a), Sc [133] (b), La [137] (c), Sm [137] (d), Eu [137] (e), Gd [137] (f).
Preprints 115443 g007
Figure 8. Self-diffusion coefficient estimated from the electrical conductivity data according to Nernst – Einstein equation for Y2W3O12 [132] (a) and Sc2W3O12 [136,141] (b).
Figure 8. Self-diffusion coefficient estimated from the electrical conductivity data according to Nernst – Einstein equation for Y2W3O12 [132] (a) and Sc2W3O12 [136,141] (b).
Preprints 115443 g008
Figure 12. Arrhenius plots of DC electrical conductivity σDC of Bi1.9Te0.1RO4.05Cl (R = Nd, Sm, Eu, Gd, Dy, Ho, Er, Tm, Yb, Lu) and Bi6–2xTe2xO8+xBr2 (x = 0.1, 0.5) in dry nitrogen. At any temperature, Bi1.9Te0.1LuO4.05Cl exhibited the highest σDC among these materials (e.g., 5.1 × 10–2 S cm–1 at 707 °C) [156]. Reprinted from Ref. [156] under the CC-BY 4.0 license.
Figure 12. Arrhenius plots of DC electrical conductivity σDC of Bi1.9Te0.1RO4.05Cl (R = Nd, Sm, Eu, Gd, Dy, Ho, Er, Tm, Yb, Lu) and Bi6–2xTe2xO8+xBr2 (x = 0.1, 0.5) in dry nitrogen. At any temperature, Bi1.9Te0.1LuO4.05Cl exhibited the highest σDC among these materials (e.g., 5.1 × 10–2 S cm–1 at 707 °C) [156]. Reprinted from Ref. [156] under the CC-BY 4.0 license.
Preprints 115443 g012
Figure 13. Ca3Co4O9+δ (CCO) structure consisting of alternating CoO2 layers and Ca2CoO3–δ rock salt layers [160]. Used with permission of Royal Society of Chemistry, from Journal of Materials Chemistry. A, Materials for Energy and Sustainability, Oxygen transport kinetics of the misfit layered oxide Ca3Co4O9+δ, Thoréton V. et al, volume 2, issue 46, Copyright (2014); permission conveyed through Copyright Clearance Center, Inc.
Figure 13. Ca3Co4O9+δ (CCO) structure consisting of alternating CoO2 layers and Ca2CoO3–δ rock salt layers [160]. Used with permission of Royal Society of Chemistry, from Journal of Materials Chemistry. A, Materials for Energy and Sustainability, Oxygen transport kinetics of the misfit layered oxide Ca3Co4O9+δ, Thoréton V. et al, volume 2, issue 46, Copyright (2014); permission conveyed through Copyright Clearance Center, Inc.
Preprints 115443 g013
Figure 15. Total conductivity of CCO-based materials: (a) Ca3Co4O9 [158], (b) Ca0.85Na0.15Co4O9 [181], (c) Ca3Co3.72Ti0.2O9 [161], (d) Ca2.9Sr0.1Co3.92O9 [161], (e) Ca2.9Ag0.4Co4O9 [178], (f) Ca2.9Bi0.1Co4O9 [202], (g) Ca3Co3.9Cu0.1O9 [193], (h) Ca2.7La0.3Co3.8Cu0.2O9 [200], (i) Ca2.85Na0.05Co3.975Mo0.025O9 [197].
Figure 15. Total conductivity of CCO-based materials: (a) Ca3Co4O9 [158], (b) Ca0.85Na0.15Co4O9 [181], (c) Ca3Co3.72Ti0.2O9 [161], (d) Ca2.9Sr0.1Co3.92O9 [161], (e) Ca2.9Ag0.4Co4O9 [178], (f) Ca2.9Bi0.1Co4O9 [202], (g) Ca3Co3.9Cu0.1O9 [193], (h) Ca2.7La0.3Co3.8Cu0.2O9 [200], (i) Ca2.85Na0.05Co3.975Mo0.025O9 [197].
Preprints 115443 g015
Figure 16. (a) Nonstoichiometry δ in Ca 2 CoO 7 q 1 + δ 2 q q CoO 2 and material’s stability at various temperatures and P O 2 values [175], (b) Phase stability diagram of the misfit CCO compound in relation to SOFC/SOEC operation modes [206]. (a) Reprinted from Journal of Solid State Chemistry, Vol. 313, Hohan Bae et al., Nonideal defect structure and high-temperature transport properties of misfit-layered cobalt oxide, Pages No. 123299, Copyright (2022), with permission from Elsevier. (b) Reprinted from Journal of Electrochimica Acta, Vol. 373, Francisco J.A. Loureiro et al., Polarisation mechanism of the misfit Ca-cobaltite electrode for reversible solid oxide cells, Pages No. 137928, Copyright (2021), with permission from Elsevier.
Figure 16. (a) Nonstoichiometry δ in Ca 2 CoO 7 q 1 + δ 2 q q CoO 2 and material’s stability at various temperatures and P O 2 values [175], (b) Phase stability diagram of the misfit CCO compound in relation to SOFC/SOEC operation modes [206]. (a) Reprinted from Journal of Solid State Chemistry, Vol. 313, Hohan Bae et al., Nonideal defect structure and high-temperature transport properties of misfit-layered cobalt oxide, Pages No. 123299, Copyright (2022), with permission from Elsevier. (b) Reprinted from Journal of Electrochimica Acta, Vol. 373, Francisco J.A. Loureiro et al., Polarisation mechanism of the misfit Ca-cobaltite electrode for reversible solid oxide cells, Pages No. 137928, Copyright (2021), with permission from Elsevier.
Preprints 115443 g016
Figure 17. Arrhenius plots for oxygen tracer diffusion coefficient (D*) and surface exchange constant (k*) for the Ca3Co4O9 (a, a’), Ca2.7Sr0.3Co4O9 (b, b’) and Ca2.4Sr0.6Co4O9 (c, c’) samples [160,207].
Figure 17. Arrhenius plots for oxygen tracer diffusion coefficient (D*) and surface exchange constant (k*) for the Ca3Co4O9 (a, a’), Ca2.7Sr0.3Co4O9 (b, b’) and Ca2.4Sr0.6Co4O9 (c, c’) samples [160,207].
Preprints 115443 g017
Figure 18. Evolution of CCO-based catalyst composition in the CLMC – SESRG – regeneration cycle [37,38,213].
Figure 18. Evolution of CCO-based catalyst composition in the CLMC – SESRG – regeneration cycle [37,38,213].
Preprints 115443 g018
Figure 22. Arrhenius plots for oxygen tracer diffusion coefficient (D*) and surface exchange constant (k*) for the Ba2Co9O14 sample [207].
Figure 22. Arrhenius plots for oxygen tracer diffusion coefficient (D*) and surface exchange constant (k*) for the Ba2Co9O14 sample [207].
Preprints 115443 g022
Figure 23. (a) General view of the LnMgAl11O19 (space group P63/mmc) crystal structure projection along the plane. (b) LnO12 tetrakaidecahedron and mirror plane containing large Ln3+ cation. (c) Ion distribution in the mirror plane of a magnetoplumbite-like structure (z = 0.75) [225]. Reprinted from Journal of the European Ceramic Society, Vol. 35, Lu Haoran et al., Thermo-physical properties of rare-earth hexaaluminates LnMgAl11O19 (Ln: La, Pr, Nd, Sm, Eu and Gd) magnetoplumbite for advanced thermal barrier coatings, Pages No. 1297-1306, Copyright (2015), with permission from Elsevier.
Figure 23. (a) General view of the LnMgAl11O19 (space group P63/mmc) crystal structure projection along the plane. (b) LnO12 tetrakaidecahedron and mirror plane containing large Ln3+ cation. (c) Ion distribution in the mirror plane of a magnetoplumbite-like structure (z = 0.75) [225]. Reprinted from Journal of the European Ceramic Society, Vol. 35, Lu Haoran et al., Thermo-physical properties of rare-earth hexaaluminates LnMgAl11O19 (Ln: La, Pr, Nd, Sm, Eu and Gd) magnetoplumbite for advanced thermal barrier coatings, Pages No. 1297-1306, Copyright (2015), with permission from Elsevier.
Preprints 115443 g023
Figure 24. Hexagonal structures of SrFe12O19 (a) and Sr0.90Gd0.10Fe11.90Cr0.10O19 (b) samples [228]. Reprinted from Journal of Rare Earths, Vol. n/a, Yuzheng Lu et al., Synergistic effects and electrocatalytic insight of single-phase hexagonal structure as low-temperature solid oxide fuel cell cathode, Pages No. n/a, Copyright (2024), with permission from Elsevier.
Figure 24. Hexagonal structures of SrFe12O19 (a) and Sr0.90Gd0.10Fe11.90Cr0.10O19 (b) samples [228]. Reprinted from Journal of Rare Earths, Vol. n/a, Yuzheng Lu et al., Synergistic effects and electrocatalytic insight of single-phase hexagonal structure as low-temperature solid oxide fuel cell cathode, Pages No. n/a, Copyright (2024), with permission from Elsevier.
Preprints 115443 g024
Figure 26. A3BC3D2O12 langasite structure viewed along the stacking axis (a) and perpendicular to the stacking axis (b) [253]. Reprinted from Ref. [253] under the CC-BY license.
Figure 26. A3BC3D2O12 langasite structure viewed along the stacking axis (a) and perpendicular to the stacking axis (b) [253]. Reprinted from Ref. [253] under the CC-BY license.
Preprints 115443 g026
Figure 27. Arrhenius plots of the total conductivity (empty symbols) and bulk conductivity (filled symbols) of La3Ga5–xGexO14+x/2 for x = 0 – 0.5 [253]. Reprinted from Ref. [253] under the CC-BY license.
Figure 27. Arrhenius plots of the total conductivity (empty symbols) and bulk conductivity (filled symbols) of La3Ga5–xGexO14+x/2 for x = 0 – 0.5 [253]. Reprinted from Ref. [253] under the CC-BY license.
Preprints 115443 g027
Figure 28. Atomistic diffusion paths of protons around O1 (black), O2 (blue) and O3 ions (red) in langasite as determined by density functional theory calculations [260]. Reprinted from Solid State Ionics, Vol. 278, Tor Svendsen Bjørheim et al., Protons in piezoelectric langatate; La3Ga5.5Ta0.5O14, Pages No. 275-280, Copyright (2015), with permission from Elsevier.
Figure 28. Atomistic diffusion paths of protons around O1 (black), O2 (blue) and O3 ions (red) in langasite as determined by density functional theory calculations [260]. Reprinted from Solid State Ionics, Vol. 278, Tor Svendsen Bjørheim et al., Protons in piezoelectric langatate; La3Ga5.5Ta0.5O14, Pages No. 275-280, Copyright (2015), with permission from Elsevier.
Preprints 115443 g028
Figure 29. Layered swedenborgite structured YBaCo4O7+δ-based compounds. Triangular and Kagomé layers are in the geometrical arrangement of the CoO4 tetrahedra within a given ab plane [265]. Used with permission of Royal Society of Chemistry, from Journal of Materials Chemistry. A, Activity of layered swedenborgite structured Y0.8Er0.2BaCo3.2Ga0.8O7+δ for oxygen electrode reactions in at intermediate temperature reversible ceramic cells, Ji-Seop Shin et al, volume 9, Copyright (2021); permission conveyed through Copyright Clearance Center, Inc.
Figure 29. Layered swedenborgite structured YBaCo4O7+δ-based compounds. Triangular and Kagomé layers are in the geometrical arrangement of the CoO4 tetrahedra within a given ab plane [265]. Used with permission of Royal Society of Chemistry, from Journal of Materials Chemistry. A, Activity of layered swedenborgite structured Y0.8Er0.2BaCo3.2Ga0.8O7+δ for oxygen electrode reactions in at intermediate temperature reversible ceramic cells, Ji-Seop Shin et al, volume 9, Copyright (2021); permission conveyed through Copyright Clearance Center, Inc.
Preprints 115443 g029
Figure 30. Temperature dependencies of total conductivity for Y0.5Ca0.5BaCo3ZnO7+δ (a) [273], (Y0.5Ca0.5)0.7In0.3BaCo3ZnO7+δ (b) [273], Y0.5Ca0.5BaCo3.2Ga0.8O7+δ (c) [274] and Y0.5Ca0.4In0.1BaCo3.2Ga0.8O7+δ (d) [274].
Figure 30. Temperature dependencies of total conductivity for Y0.5Ca0.5BaCo3ZnO7+δ (a) [273], (Y0.5Ca0.5)0.7In0.3BaCo3ZnO7+δ (b) [273], Y0.5Ca0.5BaCo3.2Ga0.8O7+δ (c) [274] and Y0.5Ca0.4In0.1BaCo3.2Ga0.8O7+δ (d) [274].
Preprints 115443 g030
Table 1. Catalytic activity of various A2Mo3O12-based catalysts in partial oxidation of methane at 750 °C and 1 atm [145].
Table 1. Catalytic activity of various A2Mo3O12-based catalysts in partial oxidation of methane at 750 °C and 1 atm [145].
Catalyst CH4 concersion,
[%]
Selectivity, [%]
HCHO CO CO2
Al2Mo3O12 2.4 21.4 71.6 7.6
Ga2Mo3O12 4.1 18.8 66.2 15.0
In2Mo3O12 10.5 10.1 65.5 24.4
Sc2Mo3O12 8.2 5.2 52.0 40.2
Cr2Mo3O12 11.7 9.2 77.2 13.3
Fe2Mo3O12 7.7 30.3 63.3 6.4
Li/Fe2Mo3O12 8.0 35.0 54.5 8.4
Zn/Fe2Mo3O12 13.6 25.0 67.2 7.9
Ce/Fe2Mo3O12 15.2 22.2 69.5 8.3
Fe/Fe2Mo3O12 20.2 13.2 73.0 13.8
Table 2. Values of oxygen vacancy and hydration energy, energy barriers for rotating and hopping steps of proton migration for CCO and Fe-doped CCO obtained from DFT calculations [189].
Table 2. Values of oxygen vacancy and hydration energy, energy barriers for rotating and hopping steps of proton migration for CCO and Fe-doped CCO obtained from DFT calculations [189].
Ca3Co4O9 Ca3Co3.8Fe0.2O9
V O formation energy, [eV] 0.82 0.65
Hydration energy, [eV] –0.11 –0.21
Rotate energy barrier, [eV] 0.06 0.05
Hopping energy barrier, [eV] 1.28 1.02
Table 3. Chemical diffusion coefficient (Dchem) and surface exchange constant (kchem) values at 600 °C for Ca3Co4O9-based materials [189].
Table 3. Chemical diffusion coefficient (Dchem) and surface exchange constant (kchem) values at 600 °C for Ca3Co4O9-based materials [189].
Sample Switch Dchem, [cm2 s–1] kchem, [cm2 s–1]
Ca3Co4O9 Dry air → 50 % O2 7.94 × 10–5 8.11 × 10–4
Dry air → wet air 6.65 × 10–5 7.45 × 10–4
Ca3Co3.8Fe0.2O9 Dry air → 50 % O2 2.34 × 10–4 1.11 × 10–3
Dry air → wet air 1.50 × 10–4 1.01 × 10–3
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.
Copyright: This open access article is published under a Creative Commons CC BY 4.0 license, which permit the free download, distribution, and reuse, provided that the author and preprint are cited in any reuse.
Prerpints.org logo

Preprints.org is a free preprint server supported by MDPI in Basel, Switzerland.

Subscribe

© 2024 MDPI (Basel, Switzerland) unless otherwise stated