Next Article in Journal
Homogenization of Radial Temperature by a Tungsten Sink in Sublimation Growth of 45 mm AlN Single Crystal
Next Article in Special Issue
Effect of Particle Size and Crystal Surface of CeO2 on the Catalytic Combustion of Benzene
Previous Article in Journal
Striped Non-Uniform Corrosion Behavior of Non-Equiatomic FeMnCoCr High-Entropy Alloy Prepared by Laser Melting Deposition in 0.1 M H2SO4 Solution
Previous Article in Special Issue
Enhancement of Titania Photoanode Performance by Sandwiching Copper between Two Titania Layers
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Complex Catalytic Materials Based on the Perovskite-Type Structure for Energy and Environmental Applications

by
Florin Andrei
1,2,
Rodica Zăvoianu
1,3,* and
Ioan-Cezar Marcu
1,3,*
1
Laboratory of Chemical Technology & Catalysis, Department of Organic Chemistry, Biochemistry & Catalysis, Faculty of Chemistry, University of Bucharest, 4-12, Blv. Regina Elisabeta, 030018 Bucharest, Romania
2
Interdisciplinary Innovation Center of Photonics and Plasma for Eco-Nano Technologies and Advanced Materials, National Institute for Laser, Plasma and Radiation Physics, 409 Atomistilor Street, 077125 Magurele, Romania
3
Research Center for Catalysts and Catalytic Processes, Faculty of Chemistry, University of Bucharest, 4-12 Blv Regina Elisabeta, 030018 Bucharest, Romania
*
Authors to whom correspondence should be addressed.
Materials 2020, 13(23), 5555; https://doi.org/10.3390/ma13235555
Submission received: 12 November 2020 / Revised: 1 December 2020 / Accepted: 3 December 2020 / Published: 5 December 2020
(This article belongs to the Special Issue Catalysts for Energy and Environmental Applications)

Abstract

:
This review paper focuses on perovskite-type materials as (photo)catalysts for energy and environmental applications. After a short introduction and the description of the structure of inorganic and hybrid organic-inorganic perovskites, the methods of preparation of inorganic perovskites both as powders via chemical routes and as thin films via laser-based techniques are tackled with, for the first, an analysis of the influence of the preparation method on the specific surface area of the material obtained. Then, the (photo)catalytic applications of the perovskites in energy production either in the form of hydrogen via water photodecomposition or by methane combustion, and in the removal of organic pollutants from waste waters, are reviewed.

1. Introduction

The 21st century has brought new challenges to scientists in solving the issues generated, on the one hand, by the increased demand for energy and, on the other hand, by the more stringent exigencies in environmental protection. In recent years we witnessed a continuous search to find more economically viable processes to generate cleaner energy starting from well known as well as from new renewable resources using more performant catalysts than conventional ones. In this respect, researchers have been focused on developing new catalysts and to perfect the existing ones in order to enhance the sustainability of energy-generating processes. An example illustrating the development of the research trends concerns the utilization of perovskite-type materials as catalysts for different processes involving energy production as well as removal of pollutants by oxidation processes. Complex oxides having an ABO3 perovskite-type structure are possible candidates for catalytic combustion, their potential as oxidation catalysts being studied for the first time approximately 50 years ago [1,2]. These oxide materials are very active, highly thermally stable and present low volatility. Due to these properties, they were considered as potential candidates to replace the expensive oxide-supported noble metals (especially platinum and palladium) which are currently the combustion catalysts used on a large scale. The utilization of perovskite-type materials for this aim allows surpassing the issues posed by the formation of noble metals volatile oxides or by their sintering at high temperatures [3,4,5].
Photocatalysis is considered the key of solving energy generation and environmental pollution problems since it can use sunlight, which is one of the cleanest energy sources. Perovskites are intensively studied as photocatalysts, for several reactions such as: generation of hydrogen by photodecomposition of water and total photo-oxidation of different organic pollutants [6,7].
The purpose of this literature review is to highlight the state of the art concerning the influence of the physico-chemical properties of perovskite-type materials on their catalytic and photocatalytic efficiencies, focusing mainly on photoelectrochemical water splitting, the photodecomposition of organic pollutants and the catalytic combustion of methane reactions. To this aim, the scientific information from more than 300 references published during the last 20 years has been screened.

2. General Aspects Concerning Perovskites Materials

2.1. Inorganic Perovskites

Most of the chemical compounds having the general formula ABO3 with r(An+) ≈ 2r(Bm+) (where r is the ionic radius of A and B cations with different valences forming compounds such as: AIBVO3, AIIBIVO3 and AIIIBIIIO3) show an inorganic perovskite-type structure. Perovskites’ name derives from the first found mineral CaTiO3 which has the same atomic arrangement [8]. The unit cell of CaTiO3 is represented as a cube, where the corners are occupied by calcium ions, the titanium ions are positioned at the body’s center, and oxygen ions at the faces center (Figure 1). The name and the simple cubic structure were preserved even though in 1946 Megaw determined that CaTiO3 has an orthorhombic structure [9]. In the perovskite structure, the larger cation A is 12-fold coordinated with oxygen ions and responsible for the stability properties, having minor effects on the catalytic properties. Generally, the cations A are elements with inert d0 and f0 electronic structure, such as alkaline or rare earth cations. The catalytic activity is influenced by the smaller B cations with an octahedral coordination, which are usually 3d, 4d or 5d transition metals acting as active sites, because they have the ability to perform redox cycles without structure destruction [8,10].
In 1926, Goldschmidt [11] has defined a tolerance factor:
t = (rA + rO)/21/2(rB + rO)
where rA, rB and rO are the ionic radii of A and B cations and O anions, respectively, which is correlated to the thermodynamic and structural stability. The ideal perovskite cubic structure is stable only if 0.8 < t < 0.9, this range being slightly larger in the case of distorted perovskite structures. At room temperature, a distortion in the structure may exist, but at high temperatures, the cubic structure is formed. Besides the ideal cubic structure, other well-known symmetries for perovskites are: orthorhombic, rhombohedral, monoclinic and triclinic [12]. In catalysis, the perovskite stability during the catalytic cycles is an essential factor and it depends on the stability of the structural lattice, the valence stability of the transition metals in the reaction environment and, last but not least, the capability of defects’ formation [8]. Perovskites have the ability to incorporate mixed valance cations in their structure by either isostructural substitution of cations in a mother structure or by formation of structural anion vacancies, affecting in this way the structural and catalytic properties [8]. The most used substitutions involve the replacing of half of B cations with other cations having different charge (ABx+0.5By+0.5O3). This type of replacement determines a shift of the oxygen ions present in the ordered structure toward the cations with higher valence. Moreover, the A and B cations in perovskites can both be easily substituted leading to doped compounds with the formula A1-xA’xB1-yB’yO3 [12]. Perovskite materials also show excellent properties for several applications, such as: dielectric, piezoelectric, semiconductors, electro-optic or superconductors [13]. Their utilisation as catalysts for energy production and environmental protection is detailed in Section 5 of this review.

2.2. Hybrid Organic-Inorganic Perovskites

Hybrid organic-inorganic perovskites, denoted HOIPs, are a class of materials derived from inorganic perovskites with general formula ABX3. Compared to the pure inorganic perovskite, in HOIPs the A site and/or X site are substituted by cations of organic amines and different inorganic/organic anions X, respectively. In this way, the rigidity and the compactness of the perovskite structure are diminished, and the organic part offers more functionalities and enhanced flexibility [14,15,16]. Several examples of A and X organic ions that can be integrated in a hybrid organic-inorganic perovskite structure are presented in Figure 2.
In 1978, the German researcher Dieter Weber reported for the first time the cubic phase of a hybrid perovskite (MA)PbX3, where MA is metylammonium and X is a halogen anion (Cl, Br or I) [16]. This structure shows an octahedral coordination around Pb2+ cations with halogen anions at the center of the faces of the cube, and the metylammonium cations occupy the A sites at the cube corners. Similar to the inorganic perovskites, the structure of HOIPs is defined by a Goldschmidt tolerance factor (t). The replacement of A and X sites in HOIPs with organic linkers leads to an adaption of this factor generating a formula with higher complexity, as follows:
t = (rAeff + rXeff)/21/2(rB + hXeff/2)
where rB is the ionic radius of B cation, rAeff is the effective radius of A cation and rXeff, hXeff are the effective radius and the height of X anion modeled as a rigid cylinder, respectively. Many HOIPs show a tolerance factor ranging in the interval 0.8 < t < 1 [17].
The HOIPs are greatly utilized with excellent results for energy production as photovoltaics, while their efficiencies in catalytic applications are very low due to their instability. Considering the highlight of this scientific work and taking into account their instability in the reaction environment, this review will be focused only on inorganic perovskite materials.

3. Preparation of Inorganic Perovskite Materials

3.1. Preparation of Powders via Chemical Routes

3.1.1. Co-Precipitation Method

One of the simplest and utilized synthesis methods for perovskite catalysts is co-precipitation. This method implies the precipitation of the metal precursors (oxides, alkoxides, nitrates or other inorganic salts), when the solubility limit decreases as an effect of adding a chemical reagent. Nitric acid can be used for dissolving the metal precursors having a low solubility [18,19], while an ammonium hydroxide solution can be used to adjust the pH in order to facilitate the precipitation [18,19,20]. The precipitate obtained is further aged, filtered and washed with deionized water until the salts in excess are completely eliminated. Finally, the resulting precipitate cake is dried, calcined and activated [21]. In order to obtain homogeneous products, it is necessary to precisely control some parameters, such as temperature, pH, coprecipitation rate and the concentrations of precursors. The main advantage of using this preparation method is that the resulting materials present higher specific surface area (SSA) than those prepared by other methods [22,23].

3.1.2. Synthesis from Amorphous Precursors—“Citrate” Method

The “Citrate” method is a preparation technique which offers an extremely good control of the stoichiometry of reaction components. Moreover, it shows high reproducibility and an enhanced degree of homogeneity of the reaction mixture [24,25]. The principle of this method is based on a complexation reaction between metal cations (which were previously added in solution by dissolving specific metal nitrates in deionized water) and a chelating agent (generally citric acid, or ethylene glycol). The molar ratio between metal ions and complexing agent is 1:1 [26,27,28]. Other substances that can be used as chelating agents are: ethylenediaminetetraacetic acid (EDTA) [29], oxalic acid [30], tartaric acid [31] or glucose [32]. The resulting mixture is heated at 80–90 °C leading to a very viscous solution. The complex formed between metal cations and organic ligand is finally calcined to obtain the mixed oxide [26,27,31].

3.1.3. Combustion Synthesis Method

In this method, stoichiometric amounts of metal nitrates of the desired cations are brought in aqueous medium together with urea (or glycine, citric acid, glucose), which is used as sacrificial fuel. Nitrate salts are used not only for the strong oxidizing character of NO3 anions, but also for their high solubility in water. Urea is the most used organic fuel especially due to its low price [33,34]. Biamino and Badini have demonstrated on LaCrO3 that this process can be divided in two steps [35]. The first step is, in fact, the synthesis of the perovskite, which is an endothermic reaction, while the second step is the exothermic reaction between oxygen from nitrates and urea. Since the endothermic reaction, implying the transformation of nitrates into the desired oxides, requires a high amount of energy to be completed, the necessary energy is provided by the oxidation of urea. Finally, a stabilization treatment is applied and the catalysts are calcined for 4 h in air in order to remove all carbonaceous deposits [36]. This method can be used to form perovskite oxides having a nanometric particle size.

3.1.4. Hydrothermal Synthesis

When applying this method, the metallic precursors are dissolved in water or are brought in the form of a slurry under high pressure and temperature conditions. An important advantage of this technique is that crystalline powders can be obtained without calcination. The particle size and shape can be modified by controlling the reaction temperature, pH and the reagent’s concentration [37,38,39].

3.1.5. Solid State Reactions

By this method, the metallic precursors (generally nitrates, carbonates or oxides) can be mixed with enhanced accuracy of the stoichiometric ratios of cations. Practically, the reagents are ball milled in a milling container. The resulting material is dried at 100 °C and calcined at 600 °C for 4–8 h in air. Afterward, the product is ground and sieved in order to collect solid particles with similar sizes and further calcined for longer time (5–15 h) at higher temperature (1300–1600 °C). The final product is again ground and sieved to collect the granular fraction having the appropriate size for its further utilisation [40,41,42].

3.1.6. Influence of the Preparation Method on the Specific Surface Area of the Perovskite-Type Materials

It is well known that the specific surface area (SSA) plays a crucial role in the catalytic processes. Depending on the preparation method of the catalysts, the porosity, size and shape of pores and the pore distribution are significantly different. An optimization of previously mentioned structural parameters can lead to the design of highly active catalysts [43]. Next, different classes of materials based on perovskite-type structure are presented taking into account their textural characteristics and their ability to be an efficient candidate for catalytic reactions.
The influence of the preparation method on the specific surface area of undoped and doped ABO3 perovskites with A = Ba, Y, Pb, La, Dy and B = Al, Cr, Ni, Cu, Ru, Ce, is resumed in Table 1. Barium and lead titanates prepared by a solid state reaction show a very small SSA (<1 m2/g) [44]. Similarly, pure LaCuO3 prepared by the same method show a very small SSA of about 0.6 m2/g [45]. A slight increase of the SSA is observed for neat LaCrO3 and LaNiO3, but it does not exceed 5 m2/g [45]. The high values of the calcination temperatures involved in the preparation of perovskites via solid state reaction are responsible for the small values of the SSA. The utilisation of high temperatures during the calcination procedure leads to non-porous materials with a non-uniformity of the particle shape and size, and with small SSA [46]. Undoped LaCrO3 and doped LaCrO3 with different contents of Mg were prepared by the citrate method. The generated SSA are similar in the range of 5–7 m2/g [47]. The increase of the SSA of pristine LaCrO3 prepared by citrate method compared to the one prepared by solid state reaction (see Table 1) can be correlated with the calcination temperature, which is much lower compared to the temperature used for the preparation via solid state reaction. The addition of MgO to the perovskite powder leads to an increase of the SSA up to 37 m2/g [48]. SSAs in the range of 4–21 m2/g were obtained by the citrate method for Fe and Co doped LaCuO3 perovskites [49,50]. Pure LaNiO3 obtained by the plasmochemical method, which is based on the injection of metallic precursors into a reactor with an air plasma having a temperature in the range of 4000–6000 K, shows an SSA of 17 m2/g [51]. The freeze-drying technique was used to prepare La1-xSrxB1-yNiy with SSA in the range of 10–16 m2/g [52,53]. In this method, the solvent is directly sublimed from the solid ice into vapor, avoiding the formation of the liquid phase which can alter the morphological and chemical homogeneity of the final product [54].
In Table 2 the influence of the preparation method on the specific surface area of lanthanide cobaltate-based perovskites is resumed. Generally, the co-precipitation method is used for both pure and doped cobaltates having rare-earth elements in the A site. The specific surface area of perovskites prepared by this method ranges from 1.6 to 8 m2/g [3,56,57]. The calcination temperature of the cake (precipitate), the calcination time and the heating rate are experimental factors controlling the specific surface area. Similar values of SSA were obtained by using the same preparation method for PrCoO3, NdCoO3 and GdCoO3, respectively [56]. The citrate method is a suitable technique for cobaltate perovskites with higher specific surface area. It is frequently used for both pure and doped complex perovskites. The specific surface areas are extended over a larger range starting from 6 m2/g for pure LaCoO3 and reaching a value of 18 m2/g for Ce-doped LaCoO3 [50,58,59]. Solid state reaction offers a smaller range of specific surface areas with a maximum value of 5.1 m2/g for Ba-doped LaCoO3 (20% of A site). Moreover, a slight increase of SSA with the increase of the ionic radii of dopants (Ca2+, Sr2+, Ba2+) is observed [45]. Several microstructured powders of different perovskite-type materials based on the La1-xSrxCo1-yB’yO3 formula prepared by freeze-drying method were reported to show SSA ranging from 10.4 to 22.7 m2/g. Yamazoe et al. reported that the freeze-drying method allows the formation of higher surface area materials compared to other thermal evaporation techniques due to its great control of evaporation processes and lower decomposition temperatures [60]. A much higher value of SSA was reported for La0.9Ce0.1CoO3 prepared by flame pyrolysis (62 m2/g) compared to the same perovskite prepared by the citrate method (10 m2/g) [58,61]. This method implies the combustion in the flame of a solution containing the metallic precursors. It is a versatile technique which ensures an excellent control of the material crystallinity and particle size through its experimental parameters [62].
Manganates represent a class of perovskite materials having manganese in B site positions (AMnO3). The effect of the synthesis method on the SSA of lanthanide manganate-based perovskites is presented in Table 3. Compared to cobaltates, the SSA of manganates prepared by co-precipitation are higher with a maximum value of ca. 15 m2/g for pure LaMnO3 [3]. For manganese-based perovskites, one of the highest values of the SSA (68 m2/g) was obtained by the citrate method, which consisted of adding citric acid and ethylene glycol to a solution containing lanthanum and manganese nitrates. It has been observed that the specific surface area increases with the citric acid etching time [66]. A detailed study concerning both the neat and Mn-doped LaAlO3 perovskites reported that SSA increases with the Mn content reaching a maximum value of 33 m2/g for LaAl0.2Mn0.8O3. When aluminum cations are completely replaced with manganese cations, SSA decreases to 22 m2/g [55]. Pure LaMnO3 prepared by flame-pyrolysis shows an SSA of 56 m2/g. By substituting (10 at.%) of La cations with Sr2+, a slight decrease of the SSA can be observed (51 m2/g). By increasing the dopant concentration (20 at.%), the SSA increases up to 70 m2/g. In contrast to Ce-doped cobaltates prepared by flame-pyrolysis (62 m2/g), La0.9Ce0.1MnO3 shows a higher SSA of about 84 m2/g, with a particle size of 30–45 nm [61]. Yu et al. reported an accurate study concerning the effect of the preparation method on the SSA of Pd-doped LaMnO3 [67] showing that an inappropriate ratio between the metallic precursors and the organic fuel in the combustion synthesis leads to a very low SSA (1 m2/g), while higher SSA can be obtained by using amorphous citrate (12 m2/g) and flame pyrolysis (32 m2/g) methods. The ultrasonic spray combustion method generates Pd-doped LaMnO3 perovskite material showing the highest SSA (39 m2/g). In contrast to the classical combustion synthesis, this method uses an ultrasonic spray gun to initiate the reaction [68]. Notably, the crystal structure of the perovskite depends on the preparation method as well. Indeed, a rhombohedral crystal structure was obtained via ultrasonic spray combustion and citrate methods, while combustion and flame pyrolysis methods generated an orthorombic perovskite structure [67]. Ciambelli et al. reported the preparation of Ce- and Y-doped rare earth manganates by the mechanochemical method. Starting from oxide and carbonate precursors, the resulting materials showed SSAs of 14 and 19 m2/g, respectively [51].
Table 4 shows the effect of the synthesis method on the SSA for lanthanide ferrite-based perovskites. The co-precipitation method leads to LaFeO3 (21 m2/g) solids having even higher SSA than manganates (15 m2/g) and cobaltates (8 m2/g) [3,57,76]. The citrate method is one of the most used preparation techniques for pure LaFeO3 and doped LaFeO3 perovskites. Pure LaFeO3 showing the SSA of 2.9 m2/g was prepared by this method. It was observed that the SSA increases with the concentration of Mg2+ used as B-site dopant. This tendency is maintained for a maximum Mg2+ concentration of 40 at.%. A further increase of the dopant content leads to the reduction of SSA [77]. A similar value of the SSA for neat LaFeO3 prepared via citrate method was reported elsewhere [78]. No changes of the SSA were observed for La0.7Ca0.3FeO3 compared to simple LaFeO3. However, when increasing the Ca2+ content to 50 at.%, the SSA dramatically decreases down to 0.7 m2/g. Other publications report values of SSA of ca. 20 m2/g for undoped LaFeO3 prepared by the citrate method. Also, it increases to ca. 38 m2/g for Ca-doped LaFeO3 [76,79]. The substantial differences between the SSA of the same perovskite materials can be correlated to the calcination temperature used during the preparation procedure. Materials with smaller SSA were calcined at 800 °C for 5 h, while the others were obtained at 700 °C in 6 h [76,79]. Pd-doped LaFeO3 shows similar values of SSA as Pd-doped LaMnO3, when prepared via citrate and combustion methods. However, the SSA decreases for Pd-doped LaFeO3 prepared by flame pyrolysis and ultrasonic spray combustion methods. In contrast to Pd-doped LaMnO3, the formation of the orthorombic crystal structure of Pd LaFeO3 is independent on the preparation method [67]. Also, the mechanochemical method is used for the preparation of doped ferrites. The resulting materials possess SSA smaller than 10 m2/g [51].
The effect of the preparation method on the SSA of supported perovkites is resumed in Table 5. The wet impregnation technique was used to stabilize commercial γ-Al2O3 (having initial SSA of 200 m2/g) with 5 wt.% La2O3. Lanthanum nitrate, manganese acetate and urea were used to load the previously prepared support with 30 % LaMnO3 by using the deposition technique. This method is based on the precipitation of the active phase, which in this case is LaMnO3 perovskite, to the support surface. Finally, the sample was dried at 120 °C and then calcined in air for 3 h at 800 °C. The resulting material shows an SSA of 88 m2/g. Yet, the calcination temperature is not enough to obtain a pure perovskite, tiny amounts of La2O3 and La(OH)3 side-phases being observed by X-ray diffraction. MgO-supported LaMnO3 with an SSA of 25 m2/g was also prepared by the same method [70]. La0.8Sr0.2MnO3 was successfully dispersed on different MAl2O4 spinels supports (M = Mg, Ni, Co) via wet impregnation technique. The perovskite loading was 20 wt.% and the SSA varies between 18 and 34 m2/g [73]. Similar SSA were obtained for Ag-doped LaMnO3 loading metallic foils made of Fe-Cr-Al using the same impregnation technique. Before the impregnation, the metallic foil was washcoated with 84.8% Al2O3, 14.4% TiO2, 0.8% La2O3 through the citrate method [80]. Neat LaFeO3 prepared via citrate method was dispersed on similar supports, leading to materials with SSA of 7.7 m2/g [81]. The highest specific surface area was obtained for ZrO2-supported LaMnO3 (132.5 m2/g) [82].

3.2. Thin Films Manufacturing Using Laser-Based Techniques

3.2.1. Pulsed Laser Deposition (PLD)

Pulsed laser deposition (PLD) is a technique belonging to the physical vapor deposition (PVD) class which takes place in a vacuum chamber. The PLD setup is schematically presented in Figure 3.
A pulsed laser beam is focused onto the surface of a target consisting of the desired material to be deposited. When the laser pulses have high enough energy density, a plasma plume forms at the target surface as an effect of vaporizing or ablating small parts from it. The flux of material necessary for the film growth is provided by the ablation plume and it is collected on a substrate. In other words, the deposition process using laser ablation is based on the vaporization of the target material, followed by the deposition of the vaporized material onto a collecting substrate. The substrate is situated at a well-known distance from the target and it is placed parallel to the target, as can be seen in Figure 4 [85,86].
The fundamental processes taking place during the laser ablation are as follows: the heating of the irradiated material; the melting, evaporation or sublimation of the heated material; the formation of the plume and, finally, the plume expansion. One of the most important ablation parameters is the energy density of the laser pulse or its fluence (J/cm2). The preparation of thin films by using PLD can be performed both in a vacuum and in a gas atmosphere, which influences the deposition process. Figure 5 displays a photograph of a TiO2 target [85,86].
Pulsed laser deposition is a non-conventional technique, relatively simple and suitable for the production of thin films that can be successfully applied for many classes of materials when other techniques fail. It presents the following advantages:
  • the laser radiation can be well focused on very small spot sizes at the target surface, increasing in this way the efficiency, the control and the flexibility of the process;
  • the deposition chamber can be considered a “clean reactor” because the energy source (laser) is external, being independent of the deposition medium; also, the laser parameters (energy density and wavelength) can be easily adjusted to ensure the reproducibility of the sample preparation;
  • it is a simple and versatile technique from the point of view of experimental achievement, offering the possibility to obtain all kind of materials (complex stoichiometry, organo-metallic compounds);
  • the properties of the obtained thin films (thickness, crystalline structure, stoichiometry and composition) can be rigorously controlled, because they depend on the laser parameters (wavelength, laser fluence, the spot area, the duration of pulse, the repetition rate etc.) which are easily controlled from the outside of the deposition chamber;
  • it ensures large deposition rates (1–5 Å/pulse).
However, the pulsed laser deposition technique has also a number of disadvantages, such as:
  • the possibility to cover only substrates having small area (~1 cm2);
  • the appearance of material droplets or clusters on the surface of the thin films leading to an increased roughness, which can affect the crystallinity, the optical, electrical and magnetic properties of the manufactured thin films [85,86].
The latter disadvantage can be diminished or even eliminated by optimizing the PLD system and deposition conditions as follows:
  • the selection of a suitable target material: a target made by dense and very small particles, ensures uniform conditions during the ablation process. A material presenting a lot of defects or different structural mechanical strains, which can appear during the processing procedure, affects the deposition process. Moreover, the target material has to present a high absorption coefficient at the used laser wavelength;
  • the rotating and the translation of the target material toward the laser beam during the deposition process;
  • the optimizing of the deposition parameters (the laser fluence, the laser spot area, the repetition rate);
  • the utilization of a supplementary laser beam parallel to the substrate surface which can split the material clusters.
Additionally, the PLD can be coupled with a radiofrequency (RF) plasma source for better performance. Compared to the standard PLD setup, in this case a RF plasma source directed to the substrate is added. The function of the RF source is to ensure a supplementary control of the anionic composition of the manufactured thin films by using gases from the RF plasma source [87]. Figure 6 shows the experimental setup of the RF-assisted PLD vacuum chamber.
In Table 6 the experimental conditions for different types of perovskite materials grown by PLD and PLD-RF techniques are presented.

3.2.2. Matrix-Assisted Pulsed Laser Evaporation (MAPLE)

A major disadvantage of the manufacturing of thin films by laser ablation arises from the deposition process. The plasma formation and the condensation of elements on a substrate are not suitable for soft organic and polymeric materials, because their structure can be decomposed easily or even completely destroyed as an effect of interaction with laser beam. In order to depose this kind of materials, a modification of the experimental setup is required, this modification being related to the deposition target. Therefore, in the matrix-assisted pulsed laser evaporation (MAPLE) technique the target is prepared by dissolving the polymeric (or organic) material into a volatile solvent and the resulting mixture is frozen in liquid nitrogen. A photograph of a MAPLE target can be seen in Figure 7. The laser wavelength is selected in such a way that only the solvent reacts when the laser beam hits the frozen target [143,144].
The difference between this technique and PLD lays in the target preparation, which generates other laser-material interaction mechanisms [145]. The setup used for MAPLE deposition is presented in Figure 8.
In this case, two main processes occur when the laser beam falls onto the target surface: the evaporation of the frozen target and the ejection of the organic material. Generally, the concentration of polymeric (or organic) substance is very low (ca. 1–5%), the evaporation of the solvent and of the desired substance taking place simultaneously. The energy of the photons absorbed by the solvents is converted into thermal energy which induces the heating and the evaporation of the target. The polymeric chains have sufficient kinetic energy to cross the distance between the target and the substrate, while the solvent molecules are eliminated from the reaction chamber by the vacuum pumps. When the experimental conditions are optimized, the polymeric (or organic) substances can be transferred from the target to the substrate without structural damage [146,147]. This method has also been used to obtain some perovskite thin films (Table 7).

4. Catalytic Applications of Perovskite-Type Materials

4.1. Energy Production

Nowadays, one of the major problems facing humanity is to fully cover the global energy demands which have severely increased in the last few years. Most energy is generated by using fossil fuels. The high energy demand on the market accelerates the consumption of these exhaustible resources based on carbon. Moreover, the intense utilization of fossil fuels causes the greenhouse effect and generates huge amounts of pollutants affecting the environmental safety and human’s health [151]. Hence, a new alternative for energy production must be urgently and efficiently implemented. One of the most promising ways to solve the current energy problem is to use solar energy. The energy provided by the sun is abundant, among the cleanest energy resources, which does not make the global warming status worse. Furthermore, solar energy is an ecological and renewable source, and these features are propelling it as a suitable candidate for the global supply. Solar energy can be converted to both electrical and chemical energies by using photovoltaic and photocatalytic concepts. However, until technologies based on these processes are suitable for industrial implementation, new alternatives in order to increase fossil fuels’ efficiency and to minimize the pollution are required. Among others, catalytic combustion of methane is one of the promising ways to increase the efficiency and to minimize pollution [152].

4.1.1. The Production of Energy in the Form of Hydrogen via Water Photodecomposition

Hydrogen is considered to be one of the most suitable options to replace the carbon-based fuels. It can be generated using different methods from various renewables (hydro, solar) and non-renewables (coal, natural gas, nuclear) sources. Among these, hydrogen can be produced from water through different processes, such as high-temperature decomposition (thermochemical water splitting) [153,154,155], electrolysis [156], photocatalysis [157] and photoelectrochemical water splitting [158]. Today, hydrogen is mainly synthesized by steam reforming of hydrocarbons (especially methane). It finds uses generally in petroleum refining [159], the production of ammonia [160] and in the metal refining industry [161]. For the future, it is intended to use hydrogen in fuel cells for high-efficiency power production systems which can directly produce electricity at low temperatures, with no emission of toxic by-products. In practice, inside a fuel cell, the hydrogen or a hydrogen-rich fuel is reacting with pure oxygen or oxygen from air, the resulting product being water [162]. As already discussed in Section 4.1, the solar energy is considered the most suitable alternative as global energy supplier and thus the water photodecomposition reaction is one of the most studied methods for chemical energy (hydrogen) production.
There are two similar processes for the photodecomposition of water: photocatalysis, which is based on a particulate system, and photoelectrocatalysis (also known as photoelectrochemistry) based on photoelectrochemical (PEC) cells. A schematic representation of the mechanisms involved in these two processes can be found in Figure 9.
Generally, in photocatalysis, the photocatalyst powders are freely suspended in a solution (as a slurry) or are fixed in a reactor bed. This is considered one of the most simple and cheapest methods for water photodecompositon which does not require transparent electrodes or directional illumination. As can be observed from Figure 9, the semiconductor photocatalyst particles are irradiated and the photons having energies higher than the material band gap are absorbed. The photons’ absorption is followed by the charge carriers formation, electrons (e) are excited to the conduction band (CB), positive holes (h+) being simultaneously created in the valence band (VB). Once formed, the charge carriers migrate to the photocatalyst surface. When they arrive to the solid photocatalyst/liquid electrolyte interface, electrons are reducing water to molecular H2, while positive holes are oxidizing it to O2. Unfortunately, only a small portion of the photogenerated charge carriers participate in the redox reaction, and most of them suffer bulk or surface recombination, as can be seen in Figure 10. In fact, the charge recombination is one of the biggest challenges of this type of process, because a high loss of the excited charge carriers leads to a decrease in the reaction yield [164,165].
In order to evaluate the performance of a photocatalyst, the concept of the apparent quantum yield (QE %) was introduced and it is described by the following formulae [166,167]:
QE (%) = (2 × number of H2 molecules/number of incident photons) × 100 (for hydrogen)
QE (%) = (4 × number of O2 molecules/number of incident photons) × 100 (for oxygen)
In addition to the high recombination of the photogenerated charge carriers, the photocatalytic system has other major disadvantages. The photogenerated hydrogen and oxygen gas are formed on the same photocatalyst surface and, hence, they are predisposed to form water via the so-called “surface back-reaction” (SBR). In order to avoid this reaction, the photocatalytic systems require the utilization of additional sacrificial reagents which are electron acceptors or donors. Moreover, the final separation of the gas produced requires additional energy consumption. Also, because the solid particles are suspended in the liquid electrolyte, a part of irradiation light will be absorbed by the liquid medium, decreasing the energy of incident light which interacts with the photocatalyst [167].
On the other hand, in a PEC system the standard configuration is based on a photoelectrode (called working electrode) and a counter electrode, as presented on Figure 9. The working electrode consists of a semiconductor material (of either n-type or p-type) deposited on a conductive substrate. When the semiconductor material is of n-type, the working electrode acts as photoanode (Figure 11a), while when it is of p-type, it operates as photocathode (Figure 11b). Moreover, the PEC system can also work as a tandem system with both photoanode and photocathode, as shown in Figure 11c.
For example, in a standard PEC configuration, when photons are absorbed by a n-type semiconductor, the positive holes from the valance band are migrating to the photoelectrode surface, where water is oxidized to O2. The electrons are collected by the conductive substrate and they are sent through an external circuit to the counter electrode (Pt), where water molecules are reduced to H2. Generally, in order to ensure a better separation of the charge carriers, an external electrical or chemical (pH difference) bias between electrodes is applied. When the working electrode is made of p-type semiconductor, O2 is formed at the counter electrode surface, while H2 is generated on the p-type photoanode surface. Also, the same principle is applied for tandem PEC configuration, water molecules are reduced to H2 at the p-type semiconductor surface while being oxidized to O2 at the surface of the n-type semiconductor [163].
The PEC system shows excellent advantages compared to the particulate system, the oxygen is generated at the photoanode, while the hydrogen is formed at the photocathode. In this way, the resulting gas can be independently collected with no further costly gas separation methods or efficiency lowering because of SBR. Furthermore, appropriate and superior quality contact between the conductive substrate and semiconductor ensures excellent charge transfer leading to great photoelectrochemical efficiencies [168].
The PEC water-splitting processes can be divided into two electrochemical redox reactions: the hydrogen evolution reaction (HER) and the oxygen evolution reaction (OER). The HER can occur by the Volmer–Heyrowsky or Volmer–Tafel mechanisms following a two-electron transfer process (2H2O + 2e ⇄ H2 + 2HO) [169]. On the other hand, the OER is a more complicated redox process implying the transfer of four electrons (2H2O ⇄ 4e + O2 + 4H+) [170,171]. Depending on the experimental conditions, as well as, the nature and the physico-chemical properties of the photoelectrodes, the backward reactions can affect the global water splitting efficiency. The hydrogen oxidation reaction (HOR) is based on the oxidation at the cathode of H2 to H+ by the same mechanism as HER. By contrast, at the anode, the oxygen reduction reaction (ORR) can directly or indirectly occur by the same four-electron transfer process. The direct reduction of O2 leads to H2O formation, and it can be associative or dissociative. On the other hand, the indirect process leads to water formation from intermediate compounds such as hydrogen peroxide [172,173]. The dissociation of H2O molecule into H2 and ½O2 is a thermodynamically non-spontaneous reaction and, hence, in order to occur, it requires an excess of energy corresponding to a Gibbs free energy of +237 kJ/mol H2 (1.23 eV/electron transfer). In other words, the water photo-dissociation can theoretically occur only if the absorbed photons have energies higher than 1.23 eV, which correspond to 1100 nm. As can be seen, the infrared light is the most suitable for water photodecompositon reaction, at least from a theoretical point of view. Nevertheless, in practice a higher energy is required because of the energy losses (electron/hole transfer, the recombination of electron/hole, kinetic losses). Generally, it is recommended to use photocatalysts/photoelectrodes having a band gap value of 2.0–2.2 eV to overcome losses. Another important factor for the overall efficiency of water splitting reaction is related to the energy position of the bands extrema. In order to make possible the HER the bottom of the conduction band (CB) of the photocatalyst must be located at a potential which is more negative than the reduction potential of H+/H2 (0 V vs. NHE—normal hydrogen electrode) and the top of the valence band (VB) has to be located at more positive potential than the oxidation potential of H2O/O2 (1.23 V vs. NHE) [174]. A schematic representation is presented in Figure 12.
In order to evaluate the performance and efficiency of materials and devices used in PEC systems, it is required to use benchmark metrics of assessment. One of the most important methods for the performance quantification of PEC systems is a photocurrent density–voltage (J–V) curve. This technique implies the measurement of the photocurrent density (mA/cm2) as a function of the applied voltage (V) under chopped or continuous irradiation. The presence of the anodic photocurrent is correlated to the O2 generation, while the cathodic photocurrent with H2 generation, respectively. Figure 13 displays the J–V curves for a photoanode in dark with/without catalyst and under illumination with/without catalyst [175]. The onset potential is defined as the applied potential where the current starts to increase; a lower onset potential is desired from an economical energy consumption point of view.
Alternatively, a direct correlation between solar energy and hydrogen evolution is necessary to calculate the solar to hydrogen (STH) efficiency. In order to determine this efficiency, it is necessary to quantitatively measure the produced H2. STH efficiency gives the overall reaction efficiency of a PEC system used for water splitting reaction under simulated solar illumination and with no external voltage applied. Also, it can be calculated only if the electrodes are immersed in the same electrolyte without other sacrificial reagents [176]. STH efficiency can be calculated using the following formula:
STH = ( v H 2 ( mmol s ) × 237,000 ( mj mmol ) P light ( mW cm 2 ) × A ( cm 2 ) ) AM 1.5 G
where, νH2 is the number of mmol of photogenerated H2, 237,000 mJ/mmol represent the Gibbs free energy, Plight is the power density of the incident light and A is the illuminated area of the electrode (area of the spot light). AM 1.5G (air mass 1.5 G) represents the standard reference solar spectra used to calculate the conversion efficiency of solar energy to electrical or chemical energy. It defines the effect produced by the atmosphere of the Earth on the solar radiation [177].
Also, STH efficiency can be directly calculated from the photocurrent density when the measurement of the H2 amount is not possible, using the formula [178]:
STH = ( J ph ( mA cm 2 ) × V redox × η f P light ( mW cm 2 ) ) AM 1.5 G
where, Jph is the photogenerated current density, Vredox is the redox potential of interest (1.23 V) and ηf is the faradaic efficiency of the hydrogen evolution reaction.
For the determination of the faradaic efficiency of hydrogen, it is crucial to know the active surface area of the catalyst (photoelectrode) and the current density passing between the two electrodes [178].
For the system where an external bias is applied, other tools are required for the calculation of the efficiency. One of them is the applied bias photon-to-current conversion efficiency (ABPE) calculated with the following formula:
ABPE = ( J ph ( mA cm 2 ) × V redox V applied × η f P light ( mW cm 2 ) ) AM 1.5 G
where Vapplied is the potential applied between the working and counter electrodes.
This formula can be applied for PEC devices working in a three-electrode system which implies the addition of a reference electrode. As in the case of STH efficiency, for the calculation of ABPE efficiency the utilization of sacrificial reagents and chemical bias should be avoided [176].
In addition, the incident photon-to-current efficiency (IPCE) is one of the most important tools for the evaluation of PEC performances. It offers the possibility to determine the practical efficiency limits of a material that can be used in PEC systems. It can be determined using the following formula [176]:
IPCE = ( J ph ( mA cm 2 ) × 1239.8 ( V × nm ) P mono ( mW cm 2 ) × λ ( n m ) )
where 1239.8 (V × nm) represents a multiplication of h (Planck’s constant) and c (the speed of light), Pmono is the power density of the monochromated light used for photoelectrode irradiation and λ is the wavelength of the light.
In 1972, Fujishima et al. reported the first photoelectrochemical water splitting reaction with solar hydrogen production opening new opportunities in the PEC field [179]. The authors used n-type TiO2 as photoanode and an irradiation source emitting in the ultraviolet (UV) region. Since Fujishima and Honda discovered this effect, TiO2 has been extensively used as photoelectrode because it is cheap, non-toxic and possesses high photostability. However, it has a wide band gap (ca. 3.2 eV; λ ≈ 396 nm) and it can absorb only the UV light which represents a small part of solar light (3–5%) [180,181].
Taking into account the excellent efficiencies of HOIPs in the photovoltaic applications, they seem to be possible candidates for integration in PEC devices. In a PEC system, the photoelectrodes must be in direct contact with an aqueous solution of electrolyte in order for the chemical reaction to take place. As already mentioned, the organic perovskites are very unstable in aqueous media, hence the presence of moisture has crucially affected their application in photoelectrochemistry [182,183]. However, many researchers have tried to overcome this drawback in order to generate a device showing excellent stability and high efficiency for PEC water splitting applications. The passivation of CH3NH3PbI3-based photoanode with a very thin Ni layer for both waterproof coating and the holes transfer has been reported. The device activity was almost entirely lost in an electrolyte solution 0.1 M Na2S (pH 12.8) after 20 min [184]. The stability was increased up to 30 min when a carbon nanotube/polymer composite layer was used as waterproof coverage layer [185]. Also, Wang et al. reported an innovative functionalization of CH3NH3PbI3-based photoanode with Ni as protective layer against humidity showing 30 min stability and 2.08 mA/cm2 value of the photogenerated current [186]. All these photoelectrodes are based on the planar perovskite solar cells, including expensive Au electrodes and hole transporting materials, making the final device very costly and irrelevant due to its unsatisfactory stability [187].
Inorganic perovskites having the ABO3 formula are of great interest for H2 and O2 production due to their exceptional stability and high photocatalytic activity. In Table 8 a series of different perovskites used as photoelectrodes for H2 and O2 generation via a photoelectrochemical water splitting reaction are presented.
Titanates were tested in photoelectrochemical water splitting in both acidic and basic environments. For example, BaTiO3 with a large band gap value (3.11 eV) shows a photocurrent density of 0.007 mA/cm2 at 0.5 V vs. SCE (saturated calomel electrode). The SCE electrode is a reference electrode based on mercury and mercury chloride [188]. The addition of PbTiO3 to TiO2 leads to a device with enhanced photoelectrochemical response (0.3 mA/cm2) compared to pure TiO2 (almost 0 mA/cm2) under visible light irradiation. The observed difference in the photocurrent density indicates that PbTiO3 acts as a visible light absorber [189]. Excellent photoelectrochemical activities are reported for SrTiO3, CaTiO3, LaNiO3 and NaTaO3 under UV radiation, but the utilization of visible light is restricted by their wide band gap values [190,191,192,193,194,195,196]. Therefore, perovskites having smaller band gap values are of interest in the photocatalytic field for enhanced water-splitting efficiencies. Ferrite perovskites are a class of materials which generally show the band gap energies in the visible region. The rhodamine B was successfully reduced under visible irradiation on YFeO3 that showed a much higher photocatalytic activity compared to TiO2 which possess a low absorption under visible light [197], and it was successfully tested in the form of both nanoparticles and compact films in photoelectrochemical production of H2 [198].
In particular, BiFeO3 (BFO) and LaFeO3 (LFO) were extensively utilized as photoelectrodes in photoelectrochemical applications, mostly for water photolysis and dye photodegradation. The dyes photodecomposition process will be detailed in the next section of this review. BFO is one of the few multiferroic materials at room temperature with a band gap value of ca. 2.7 eV and an excellent photocatalytic activity [199]. The ferroelectric properties of BFO are able to improve the electron-hole separation and, hence, the photocatalytic activity. The effect of spontaneous polarization on the charge carriers separation is extensively detailed in literature [200]. Both undoped and doped BFO prepared by PLD were tested in photoelectrochemical applications as shown in Table 8. In 0.2 M Na2SO4 electrolyte under sunlight illumination, the BFO photoanode generates a photocurrent density of 0.12 μA/cm2 at 0.8 V vs. Ag/AgCl. After the material was heated in a reductive atmosphere of hydrogen (5%), the photocurrent increases ca. 6 times. The difference is due to smaller recombination and more efficient separation of free carriers on the hydrogen-treated sample [201]. A value of 17% for incident photon-to-current efficiency was reported for BFO photoelectrodes prepared by chemical vapour deposition (CVD). The modification of BFO surface with Ni improves the reaction kinetics and decreases the overpotential for water oxidation. The photocurrent density increases from 0.17 to 0.72 mA/cm2 with Ni addition [202]. Nanowires as well as nanocubes of BFO generate only O2 when irradiated with UV light. A mixture of BFO and SrTiO3 (STO) can lead to H2 formation under visible light irradiation [203,204]. Also, in the case of polycrystalline BFO films, it was reported that by increasing the film crystallinity, the PEC response is enhanced. Moreover, the PEC response is strongly dependent on the spontaneous polarisation of BFO polycrystalline films [205,206]. Similarly, Cho et al. have reported epitaxial thin films of BFO deposited by PLD on STO showing excellent PEC properties [207]. Ti-doped BFO was reported to generate 0.04 mA/cm2 at 1.23 V vs. RHE (reversible hydrogen electrode) in a basic solution of NaOH under UV irradiation [208]. The RHE is a reference electrode with the potential independent on the pH changes. The onset potential was ca. 0.81 V vs. RHE [208]. A detailed study on Y-doped BFO used as photoanode with high PEC efficiency for water splitting reaction (Jph = 0.8 mA/cm2 at 1.4 V vs. RHE) was recently reported in literature [209]. Moreover, complex BFO-heterostructures were used as photoelectrodes for water splitting reaction. For example, nanofibers of BiFeO3/Bi2Fe4O9 with enhanced stability show higher efficiency for H2 production compared to pure BFO and Bi2Fe4O9. This improved efficiency was attributed to the formation of well-defined heterojunctions between the component materials, which helps the separation of charge carriers and avoids their recombination [210]. A significant increase in the photocurrent density compared to pure BFO was also observed for BiFeO3/Fe2O3. The heterostructure having a concentration of 9% Fe2O3 shows a Jph value of ca. 0.19 mA/cm2 (at 0.6 V vs. Ag/AgCl), using visible irradiation light, which is about three times higher than that obtained for pure BFO (0.055 mA/cm2) [211]. Recently, a complex device based on WO3/BiVO4/BiFeO3/FTO (fluorine-doped tin oxide) was reported to show a huge value of the photocurrent density under solar irradiation in 0.5 M Na2SO4 electrolyte (46.9 mA/cm2 at 2.53 V vs. RHE) [212]. The great response is correlated to the great separation of photogenerated charge carriers due to the similar band alignment of the three component materials deposited on the FTO substrate. The presence of an internal electric field at the BiVO3/BFO p-n junction interface together with BFO polarization, which improve the electron-hole pairs separation, have also been noticed [212].
LaFeO3 is a perovskite with a smaller band gap value than BiFeO3, more specifically ca. 2.07 eV, offering the possibility for a better absorption in the visible range [180,213]. Several years ago, Celorrio et al. studied the photoelectrochemical properties of nanostructured LFO photocathode manufactured by the screen-printed technique. They reported higher cathodic than anodic photocurrent values [214]. In screen-printed technology, the perovskite material is combined to an organic compound which is used to improve the adhesion to the substrate. The mixture is printed to the substrate using a heater, followed by calcination [215]. Also, in another study, the simultaneous generation of O2 and H2 using LFO powders was noted [216]. The values of the photogenerated current are around 0.8 mA/cm2 at potentials higher than 1 V vs. Ag/AgCl for utilisation of LFO as photoanode under simulated solar light [217]. Yu and coll. have deposited p-LFO and p-LFO/n-Fe2O3 on ITO (indium tin oxide) conductive substrates by PLD. The fabricated photocathodes show enhanced stability in alkaline NaOH electrolyte solution under visible irradiation even after 120 h [218]. May et al. [219] have reported an accurate study on the effect of thickness of LaFeO3/Nb:SrTiO3 (LFO/STON) ultrathin films prepared by PLD on the photoelectrochemical properties. The cathodic photocurrent was shown to be strongly dependent on the film thickness, being observed only for samples with 10–25 nm thickness. The anodic photocurrent is observed for both thinner and thicker than 10 nm LFO films [219]. A photoanode based on Cu-doped LFO was reported to show a photocurrent density of 2 mA/cm2 at 1.1 V vs. Ag/AgCl under visible irradiation [220]. Notably, no reports on utilization of devices based on heterostructures of LFO/BFO perovskites for water splitting reaction have been published to date.
The utilization of co-catalysts can further improve the global efficiency of water splitting reaction. Generally, noble metals (Pt, Rh, Pd) and metal oxides (NiO, RuO2) are extensively used as co-catalysts. When a photoelectrode is loaded with a co-catalyst, the photogenerated electrons which arrive at the surface are entrapped by the co-catalyst. In this way, the utilisation of co-catalysts, stimulates the surface chemical reaction and constrains the backward reaction, leading to an increased overall water splitting efficiency [90]. A schematic representation of a photocatalyst and a photoelectrode loaded by co-catalysts is presented in Figure 14.
Also, the working principle of the reaction in the presence of photocatalysts (a) and photoelectrodes (b) loaded with an oxidation co-catalyst for both dark and illuminated processes is illustrated in Figure 15.
Polycrystalline films of BFO synthesized via the citrate method, decorated with nanoparticles of Ag which act as co-catalysts, were reported. This heterostructure shows almost two times higher photocurrent density compared to pure BFO [229]. Also, a device based on reduced graphene oxide (rGO)/BFO was reported to ensure an excellent charge carriers separation, leading in this way to a higher IPCE [230]. Moreover, thin films of BiFeO3 show water-splitting activity without any external bias applied [200,221].

4.1.2. Catalytic Combustion of Methane

Natural gas (a hydrocarbon gas mixture consisting mainly of methane) has a major impact on our lives, being one of the most used energy sources. Thermal combustion of light hydrocarbons meets many challenges because of its low efficiency and high pollution degree. This conventional combustion process takes place at high temperatures (above 1300 °C) generating high amounts of nitrogen oxides (NOx), which severely affect the environment and, furthermore, human health. In the last few years, the market request has shown a significant increase of fuel consumption, leading, at the same time, to an advanced pollution risk. The catalytic combustion is preferred instead of the flame combustion due to its major advantages, such as enhanced fuel efficiency and lower emission of pollutants in the exhaust gas [3,5]. Important quantities of methane (CH4) are annually consumed in the fuel industry, it being considered one of the cleanest sources of fossil energy [231]. The simplified chemical reaction of methane combustion is:
CH4 + 2 O2 → CO2 + 2 H2O
However, beside the production of CO2 and H2O, during the methane combustion process, many other reactions can take place, such as partial combustion of methane with CO formation, steam reforming and others [232]:
CH4 + 3/2 O2 ⇒ CO + 2 H2O
CH4 + H2O ⇄ CO + 3 H2
2 H2 + O2 ⇒ 2 H2O
CO + H2O ⇄ CO2 + H2
2 CO + O2 ⇒ 2 CO2
In contrast to the classical combustion, the heterogeneous catalytic combustion of methane operates at lower temperatures (<800 °C), increases the conversion efficiency of methane into energy, and severely decreases the atmospheric pollutants emission [233]. It is worth noting that catalytic combustion of methane is important not only for power generation but, also, for air pollution abatement [234]. The ideal catalyst for methane combustion is required to possess high thermal and chemical stability, but at the same time it should present catalytic oxidation activity in the low-temperature range [4]. The most promising low-temperature catalysts for this reaction are the supported noble metals. Excellent catalytic activities for methane combustion are reported on supported platinum [235], rhodium [236] and palladium [237]. Unfortunately, they are expensive and in a strong oxidative atmosphere they are susceptible to form volatile oxides leading to a decrease of the catalytic activity. Also, at high temperatures, the noble metal particles tend to accumulate and to form a compact material (the so-called sintering effect). This effect decreases the specific surface area (SSA) of the catalyst and, implicitly, its catalytic activity [4]. These disadvantages made researchers look for a possible replacement of this kind of catalysts. A good alternative is to use active catalysts for hydrocarbons combustion based on single or binary oxides of transition metals. A few single oxides active for hydrocarbons combustion are Co3O4, CuO, NiO, MnO2. The combination of oxides generally gives a greater thermal stability and higher combustion activity compared to the single oxides [45].
Inorganic perovskites ABO3 have been found to be adequate catalysts for total oxidation since they combine the high catalytic activity with the low volatility [3]. For the perovskite-type oxide catalysts, two types of oxygen species having different bonding strength are present: (i) adsorbed on the surface and (ii) lattice oxygen. It is believed that the adsorbed oxygen is more active, reacting with methane at lower temperatures than the lattice oxygen. At higher temperatures, the coverage of the adsorbed oxygen decreases, while the lattice oxygen becomes highly active. Voorhoeve et al. have proposed two reaction types for catalytic combustion of methane over perovskite materials: suprafacial and intrafacial. The suprafacial reaction is mainly present at low temperature, while with increasing the temperature, the intrafacial process occurs [45].
Table 9 presents the performances of different undoped perovskites used as catalysts for methane combustion. Lanthanum cobaltate (LaCoO3), having a SSA smaller than 15 m2/g. shows T50 (temperature corresponding to 50% methane conversion) values ranging between 525 and 709 °C [45,50,56,57,58,59,63,64] depending on the reaction conditions, i.e., methane and oxygen concentrations in the feed gas and the magnitude of the gas flow rate. With the increase of the SSA, T50 decreases, reaching a minimum temperature of 449 °C for LaCoO3 with an SSA of 43 m2/g [65]. This temperature can be further lowered by adding noble metals (e.g., Pt, Pd) to the LaCoO3 perovskite structure [61]. The replacement of La3+ with Pr3+ leads to a major increase of T50 and of the activation energy (Ea), which is, in fact, the required energy for a chemical reaction to take place. This behavior is correlated to the redox properties of Pr4+/Pr3+ couple which can affect the valence state of cobalt ions during methane combustion reaction [56]. Lanthanum manganate (LaMnO3) shows excellent catalytic activity for the methane combustion with T50 in the 500–600 °C interval [50,55,57,64,70,72,75,80,82,238]. However, the value of T50 can be further decreased by changing the synthesis method of LaMnO3 perovskites. For example, LaMnO3 prepared by the flame-hydrolysis method shows a T50 of 489 °C, while that prepared by the citric acid and flame pyrolysis methods show T50 values of 446 and 435 °C, respectively [61,65,66]. No major changes in the catalytic activity were observed when La3+ was replaced by Pr3+ or Gd3+ [69]. However, it severely decreases for NdMnO3 and SmMnO3 because of their easier reductibility [72]. It is worth noting that, due to its good thermal stability and high activity, which are further improved after dispersion on a lanthanum-stabilized alumina-coated monolith [70], LaMnO3 is an active phase of choice in high-pressure structured catalytic combustors [239,240,241,242,243]. Lanthanum ferrite (LaFeO3), which shows the highest stability compared to LaCoO3, LaNiO3 and LaMnO3, displays a catalytic activity similar to LaMnO3 [3,45,57,61,63,64,65,69,76,77,78,244]. Ciambeli et al. have reported a detailed study on ferrite perovskites (AFeO3) containing different rare earth cations (La3+, Sm3+, Nd3+) in A sites. It was demonstrated that LaFeO3 catalyst shows the best catalytic activity for methane combustion, with total selectivity to CO2, having the T50 = 528 °C and the temperature corresponding to 90% methane conversion (T90) of ca. 615 °C [77]. Spinicci et al. have revealed that La0.9FeO2.85 shows better activity compared to stoichiometric LaFeO3. The further decrease of lanthanum and oxygen contents leads to the loss of the catalytic activity [245]. The catalytic activity can be limited by the low specific surface area of the perovskite catalysts [45]. One of the most used methods to increase the SSA of catalysts based on perovskites is to disperse them on appropriate supports [246]. Indeed, it has been shown that the SSA of the unsupported BaTiO3 and PbTiO3 increased from 0.4 and 0.5 m2/g, respectively, to 193 and 175 m2/g for their alumina-supported counterparts, respectively. The increase of the SSA led to an enhanced catalytic activity for methane combustion [44,247].
The catalytic activity of the perovskites is strongly dependent on B-site cations, while A-site cations are responsible for structure stability, as already mentioned in Section 2.1. Therefore, by substituting small portions of A and/or B cations with other transition metals, both the catalytic activity and thermal/chemical stability of the final catalyst can be improved. Generally, the most used A-site dopants for perovskites are alkaline earth metals (Sr, Ca and Ba) [3,45,53,58,61,71,73,74,76,78,79] and lanthanides (Ce, Eu) [45,51,58,61,71], as can be observed in Table 10 where the performance of A-site doped perovskites in methane combustion is resumed. Arai et al. [45] reported that La0.6Sr0.4MnO3 perovskite prepared by solid state reaction exhibits similar catalytic activity to Pt/Al2O3 at low temperatures (350–550 °C). At temperatures higher than 600 °C it becomes less active than the noble metal catalyst. However, La0.6Sr0.4MnO3 was shown to be the most active catalyst compared to Ca-doped LaMnO3, Sr-doped LaFeO3, Ba-doped LaCoO3, Ce-doped LaCoO3 and Ca-doped LaCoO3 [45]. The high activity of La0.6Sr0.4MnO3 for methane combustion at low temperatures was confirmed in other studies [71,74]. This behavior was correlated with its high capability to adsorb oxygen on the surface, most probably due to the fact that the presence of the bivalent dopant determines the formation of Mn4+ species, [45,71,74]. When the temperature increases, the amount of adsorbed oxygen decreases and, hence, its catalytic activity. Calcium is the most used element for LaFeO3 doping, as can be seen in Table 10. Pecchi et al. [76] reported a detailed study concerning La1-xCaxFeO3 perovskites prepared by citrate and co-precipitation methods. For the samples prepared by co-precipitation, there are almost no changes in T50 with the calcium content, while for those prepared via the citrate method the addition of calcium minimizes T50 [76]. Similar results on Ca-doped LaFeO3 prepared by the citrate method were also reported elsewhere [79].
In Table 11 the performances of B-doped perovskites in the catalytic combustion of methane are presented. Saracco et al. studied LaCr1-xMgxO3 perovskite prepared by the citrate method and observed that its catalytic activity increases with the Mg content [47]. The opposite behaviour was observed for LaAl1-xMnxO3, whose catalytic activity decreases with the Al content. LaAl0.2Mn0.8O3 shows the best activity, while the activity of LaAl0.4Mn0.6O3 is similar to the pure LaMnO3 [55]. A high fraction of Fe4+ is observed in LaFe1-xMgxO3 perovskite and it increases with the addition of Mg content leading to lower catalytic activity at low temperatures [77]. Taguchi et al. [248] reported that Ca(Mn0.6Ti0.4)O3 shows better catalytic activity compared to undoped LaFeO3 and (La0.8Sr0.2)(Cu0.15Fe0.85)O3 perovskites. In the case of Ca(Mn0.6Ti0.4)O3 catalyst, the T50 value is ca. 580 °C, while for LaFeO3 and (La0.8Sr0.2)(Cu0.15Fe0.85)O3 it increases up to 800 and 780 °C, respectively. The significantly lower T50 value of the Ca(Mn0.6Ti0.4)O3 system was attributed to the increased content of Mn3+ cations which act as oxygen adsorption sites [78,248,249]. Moreover, even a lower T50 value was reported for Al-doped LaMnO3. Thus, for 10% Al3+ content, the T50 is ca. 520 °C and increases with the Al content. Notably, the T50 value decreases with the increase of the SSA, the highest SSA (22 m2/g) corresponding to the La(Mn0.9Al0.1)O3 system [250]. Recently, Miao et al. [251] reported a catalyst based on La(Mn,Fe)O3 with an excellent stability at 550 °C in the catalytic combustion of methane. The methane conversion was almost entirely preserved after eight combustion cycles [251].

4.2. Applications of Perovskite-Type Materials in the Removal of Pollutants from Waste Waters

In the last few years, the high extent of pollution has become one of the biggest problems facing humanity because it can severely affect human life [252]. The residual dyes emerging from different industries (pharmaceutical, textile, paper and others) are considered to be the most common water pollutants. For example, in the textile industry, every kilogram of final product generates 50–100 L of waste water [253,254,255]. Numerous techniques for water treatment have already been developed, such as adsorption [256], chemical coagulation [257], electrocoagulation [258], advanced oxidation processes [259], photocatalysis [260] and photoelectrocatalysis [261]. The adsorption method used in waste water treatments involves the adhesion of pollutants (organic dyes, in this case) to the surface of a solid material, called adsorbent. The adsorbent properties of solid materials can be explained by the force fields which govern the surface properties. Depending on the type of forces established between the adsorbent and the adsorbed molecules, the adsorption can be: physical adsorption (for weak forces) and chemisorption (chemical bonds). The main characteristics controlling the adsorption efficiency of an adsorbent are the specific surface area and porosity. The adsorption method has the advantage that it can be used for a large amount of water with low pollutants content using simple and relatively cheap operating systems [262]. The most commonly used adsorbents for the elimination of organic dyes from water are: activated carbon [263,264], alumina [265,266], bentonite [267,268] and zeolites [269]. The main drawbacks of these techniques are correlated to the adsorbents’ capacity which decreases in time and the high costs of their regeneration. Moreover, depending on the pollutant’s nature, the adsorbent could be irreversibly blocked [262]. Chemical coagulation is an extensively used method for the waste water treatment. It is based on the utilization of coagulants which have the ability to form precipitates. The pollutants are trapped in the formed precipitate which is settled, offering the possibility to easily separate the supernatant (water) from sediments. The process is strongly dependent on the pH, coagulant concentration and mixing. Depending on the nature of the pollutant, the coagulant can be either inorganic (AlCl3; FeCl3) or organic (poly diallyldimethyl-ammonium chloride; polyacrylamide). This method is able to remove high amounts of organic dye from water, but it leads to an increased cost of the process due to two drawbacks: (i) it requires high quantities of chemical coagulant, and (ii) it produces large amounts of sludge [270,271]. The electrocoagulation is similar to the chemical coagulation, in this case the coagulants being electrochemically generated. It uses low-voltage and two sacrificial iron (or aluminum) electrodes. Under an applied current voltage, at anode are generated Fe3+ (or Al3+) species, while at cathode water is reduced to H2 and hydroxides. The ions formed at the anode surface react with hydroxide groups, leading to the coagulants’ production. The resulting quantity of sludge is much lower compared to the chemical coagulation [270,272]. However, it requires often the replacement of the sacrificial electrodes to preserve a high efficiency. The operating cost of this technique increases significantly when applied in non-electrified areas, this being a common case for waste water found in nature [273]. One of the most used advanced oxidation methods for the degradation of dyes from waste water is ozonation. Ozone has a very high oxidation potential being capable to oxidize the organic dyes [259]. Unfortunately, ozone is very toxic for the human body, its strong oxidizing character can cause various diseases [274]. The photocatalytic and photoelectrochemical processes are considered to be the key solution to remove environmental pollution because they are clean methods which use renewable solar energy, their advantages have been already discussed in this review.
Rhodamine B is an important cationic xanthene dye being one of the mostly used model organic dye in photodegradation studies. Important quantities of this organic dye are coming from the textile industry contributing to environmental pollution [275,276,277,278]. Equally, methyl orange (anionic dye) and methylene blue (cationic dye) are considered harmful dyes released from textiles and printing industries and it is desired to minimise as much as possible their concentration in the environment [279,280,281,282,283,284]. Besides these, other common organic pollutants targeted in photodegradation applications are congo red (highly toxic and carcinogenic pollutant) [285], neutral red [286], phenol red [287], 4-methylphenol [288] and tetracycline [289], their chemical structures being presented in Table 12.
The photooxidation reaction of organic dyes is similar to the processes involved in the water-splitting reaction. When the semiconductor is irradiated, the photons having energies higher than its band gap energy are absorbed. The photogenerated charge carriers migrate to the photocatalyst surface. Once there, electrons react with adsorbed oxygen molecules (O2(ads)) generating initially superoxide radicals (O2) and, then, hydroperoxide radicals (HOO). On the other hand, the positive holes interact with the surface-adsorbed water molecules (H2O(ads)) producing free hydroxyl radicals (OH). When the organic compounds (dyes) are adsorbed on the photocatalyst surface, they are rapidly oxidized by the highly reactive radicals to CO2 and H2O [296]. The photodegradation of organic compounds can be schematically presented by the following series of reactions, where P = photocatalyst:
P + hν → P(h+VB) + P(eCB)
P(h+VB) + H2O(ads) → P + OH + H+
OH + organic dye → CO2 + H2O
P(eCB) + O2(ads) → P + O2
O2 + organic dye → CO2 + H2O
As in the case of hydrogen production from water, TiO2 is the most utilized photocatalyst for water depollution. It was used as photocatalyst for organic dye photooxidation in both powder and immobilized form, with good photocatalytic activity [297,298,299]. Its activity was further improved when an external bias was applied to the system. However, pure TiO2 has high rate of the charge carrier’s recombination and wide band gap value, which are the main drawbacks of this semiconductor in photocatalysis [164,300].
Structural and compositional properties of oxide perovskites are suitable for the photodegradation of organic molecules, their performances being summarised in Table 13. SrTiO3 shows excellent photocatalytic activity under UV irradiation for the degradation of rhodamine B [301,302] and methyl orange [303]. The rhodamine B was completely degraded after 1.3 h, while 3 h were necessary to completely eliminate methyl orange dye. Under visible irradiation, the degradation efficiency of SrTiO3 is lower than 50%, because of its large band gap [304].
The doping of SrTiO3 with different elements (Fe, Nb) leads to an improvement of the photocatalytic efficiency under visible light irradiation [289,304,305]. LaCoO3 shows even higher photocatalytic activity under UV irradiation for the degradation of rhodamine B, with a complete photooxidation of the organic dye after 0.8 h [305]. Its activity considerably decreases under visible irradiation [306].
The degradation of methyl orange was performed with conversions higher than 90% over BiFeO3 photocatlaysts under both UV and visible light [307]. Moreover, nanoparticles of BiFeO3 and Gd-doped BiFeO3 (Bi1-xGdxO3, where x = 0, 0.05, 0.1, 0.15) prepared by classical citrate method were tested in the photodegradation of MB. It was observed that while increasing the Gd content, the photocatalytic activity increases as well, reaching the highest conversion of the organic dye (94%) with Bi0.9Gd0.1FeO3. The increased ferromagnetic nature of Bi0.9Gd0.1FeO3 is considered responsible for its different photocatalytic behavior in MB degradation [308]. A similar degradation efficiency (94%) was observed for rhodamine B over Gd-doped BiFeO3 [309]. Congo red dye was almost completely oxidized on Ca-doped BiFeO3 ultrafine nanofibres (CaxBi1-xFeO3, where x = 0, 0.05, 0.1, 0.15) and the photocatalytic conversion of the dye increased along with the Ca content [310]. LaxBi1-xFeO3 (x = 0, 0.05, 0.1, 0.15) perovskites were used as photocatalysts for the photooxidation of the same dye, almost 80% of it being oxidized on La0.1Bi0.9FeO3 after 3 h. Further increasing the La content decreases the conversion to ca. 50% [311].
A heterostructure based on BiFeO3/CuWO4 was tested for the degradation of methyl orange (MO) and rhodamine B (Rh B). The dyes’ concentrations were spectrometrically analyzed in time, and the absorption peaks of MO and Rh B disappeared almost entirely after 120 and 75 min, respectively. The higher photocatalytic activity observed on BiFeO3/CuWO4 compared to single components was correlated with the heterojunction formation between the two oxides. The most beneficial BiFeO3:CuWO4 ratio for the organic dyes photodegradation was found to be 1:1 [312]. Methyl violet (MV) was photooxidized up to 93% on a heterostructure based on p-n heterojunction BiFeO3/TiO2 under visible irradiation as an effect of the poor recombination of charge carriers and strong absorption properties of the device [313].
Great efficiencies were reported for pure LaFeO3 used as photocatalyst under visible irradiation for the degradation of 4-methylphenol [314], rhodamine B [315,316], and methyl orange [317]. Nanocubes, nanospheres and nanorods of LaFeO3 were used as photocatalysts for Rh B oxidation under visible irradiation. No remarkable differences were observed between the nanorod and nanosphere morphologies, the organic dye conversions being similar. On the other hand, for the nanocube morphology the conversion efficiency was much lower. Irrespective of their morphology, all LFO samples showed better activities than Degusa P25 TiO2.
The higher efficiency of LFO is explained by its strong absorption properties in the visible region. Furthermore, the presence of an enhanced density of adsorbed oxygen or oxygen from surface hydroxyl groups on LFO photocatlaysts, which can act as efficient oxidizing agent, was confirmed by X-ray photoelectron spectroscopy (XPS) [318]. Wei et al. [319] reported that the doping of LaFeO3 with Mn ions leads to a perovskite structure of LaFe0.5Mn0.5O3 having higher oxygen vacancies and excellent absorption properties for visible light. All these, together with the variable valency of Mn cations contribute to the generation of a photocatalyst with improved activity for photodegradation of MO. It was also noted that the reaction mechanism is pH-dependent: the global yield decreased with the increasing of the pH value [319]. Recently, numerous scientific papers based on a heterostructure of nanosheets of graphitic carbon nitride (g-C3N4)/LFO nanoparticles were reported. High values of the photogenerated current were obtained on 15% g-C3N4/LFO (~9 μA/cm2) and 20% g-C3N4/LFO (~17 μA/cm2) compared to pristine LFO (0.04 μA/cm2). Rh B was almost completely oxidized on g-C3N4/LFO after ca. 2 h under visible light irradiation. Moreover, the photocatalytic device showed excellent stability, its activity being entirely maintained after 4 catalytic cycles [320,321]. This photocatalyst was also used with good results for Brilliant Blue (BB) degradation [322].

5. Conclusions and Perspectives

The properties and (photo)catalytic behavior of perovskite materials in water splitting, catalytic combustion of methane and photo-degradation of pollutants from water have been presented and discussed based on more than 300 relevant papers, leading to the following conclusions:
There are various synthesis methods for both powder and thin films, which determine their physicochemical properties. The specific surface area of the perovskite powders, which is a key characteristic of a solid catalyst, is strongly influenced by the preparation method used, but remains low. Indeed, the highest surface areas, mainly obtained by citrate and flame-pyrolysis methods, do not exceed several tens of m2/g. On the other hand, pulsed laser deposition is one of the most suitable preparation methods for inorganic perovskite thin films, due to its high material transfer efficiency, precise control and the great flexibility of the process. Depending on the experimental conditions, the stoichiometry of the material, as well as the thickness and the crystallinity of the films can be controlled. The most commonly used lasers for the preparation of perovskite films are those emitting in UV spectrum (193 nm, 248 nm and 355 nm). The films’ thickness starts from less than 1 nm and rises up to ca. 600 nm.
Oxide ferroelectric perovskites show excellent efficiency for the conversion of solar energy into chemical energy (H2) via water splitting. Both photocatalytic and photoelectrochemical systems are extensively studied in this application domain. The high spontaneous polarization of BiFeO3 is beneficial for a very efficient electron-hole separation. LaFeO3 presents strong absorption properties of visible light, which represents ca. 42% of the entire solar spectrum. The photoelectrodes are tested for a wide range of pH values, starting from semi-acidic to strong alkaline media. The highest photocurrent density (46.9 mA/cm2 at 2.53 VRHE) is obtained for a complex heterostructure based on WO3/BiBO4/BiFeO3. The best stability (more than 120 h) was reported in 1M NaOH for p-LaFeO3/n-Fe2O3.
Due to their good thermal stability, perovskite materials were successfully used in the catalytic combustion of methane for both power generation and methane emission abatement. Although the performance of pure perovskites is limited by their small specific surface area, their efficiency can be improved either by dispersion onto support materials possessing high surface area and thermal stability or by doping with other transition metals. Indeed, substitution in A and B sites of the perovskite structure with small amounts of other cations can improve both the stability and activity of the catalyst. Improved activity and stability can also be obtained by coating of the supported perovskite either on ceramic or metallic monoliths. The most used A-site dopants for perovskites are alkaline earth metals (Sr, Ca and Ba) and lanthanides (Ce, Eu), while for B-sites metals from the 3 and 4 periods (Mg, Al, Mn and Cu) in particular are preferred. The most active perovskites for the low-pressure methane combustion is La0.6Sr0.4MnO3 with a value of T50% of 360 °C. The high activity of this catalyst is due to its enhanced ability to adsorb oxygen on the surface.
The photodegradation of organic dyes on inorganic semiconducting perovskites showed excellent results. Their high stability under extreme chemical conditions, strong absorption properties and efficient charge separation lead to high photocatalytic activity even after several reaction cycles. Catalytic systems containing BiFeO3 perovskites as such or modified with different dopants exhibited an exceptionally high activity in the photocatalytic degradation of both anionic and cationic organic dyes.
Due to their high chemical and thermal stabilities together with their large compositional flexibility, perovskites remain (photo)catalytic materials of choice for applications in energy production and environmental protection. Obviously, new preparation procedures and the improvement of existing ones will allow in the future not only higher surface areas to be obtained but also new morphologies with consequences for their (photo)catalytic behavior.

Author Contributions

Conceptualization, R.Z. and I.-C.M.; writing—original draft preparation, F.A.; writing—review and editing, R.Z. and I.-C.M.; visualization, F.A.; supervision, I.-C.M. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Libby, W.F. Promising catalyst for auto exhaust. Science 1971, 171, 499–500. [Google Scholar] [CrossRef] [Green Version]
  2. Pedersen, L.A.; Libby, W.F. Unseparated rare earth cobalt oxides as auto exhaust catalysts. Science 1972, 176, 1355–1356. [Google Scholar] [CrossRef] [Green Version]
  3. McCartyt, J.G.; Wise, H. Perovskite catalysts for methane combustion. Catal. Today 1990, 8, 231–248. [Google Scholar] [CrossRef]
  4. Ladavos, A.K.; Pornonis, P.J. Catalytic combustion of methane on La2-xSrxNiO4-λ (x = 0.00–1.50) Perovskites prepared via the nitrate and citrate routes. J. Chem. Soc. Faraday Trans. 1992, 88, 2557–2562. [Google Scholar] [CrossRef]
  5. Li, X.; Liu, Y.; Deng, J.; Xie, S.; Zhao, X.; Zhang, Y.; Zhang, K.; Arandiyan, H.; Guo, G.; Dai, H. Enhanced catalytic performance for methane combustion of 3DOM CoFe2O4 by co-loading MnOx and Pd–Pt alloy nanoparticles. Appl. Surf. Sci. 2017, 403, 590–600. [Google Scholar] [CrossRef]
  6. Rajeshwar, K. Hydrogen generation at irradiated oxide semiconductor–solution interfaces. J. Appl. Electrochem. 2007, 37, 765–787. [Google Scholar] [CrossRef]
  7. Wang, W.; Tade, M.O.; Shao, Z. Research progress of perovskite materials in photocatalysis- and photovoltaics-related energy conversion and environmental treatment. Chem. Soc. Rev. 2015, 44, 5371–5408. [Google Scholar] [CrossRef] [PubMed]
  8. Ladavos, A.; Pomonis, P. Methane Combustion on Perovskites. In Perovskites and Related Mixed Oxides; Granger, P., Parvulescu, V.I., Kaliaguine, S., Prellier, W., Eds.; Wiley: Weinheim, Germany, 2016; pp. 367–388. [Google Scholar]
  9. Megaw, H.D. Crystal structure of double oxides of the perovskite type. Proc. Phys. Soc. 1946, 58, 340. [Google Scholar] [CrossRef] [Green Version]
  10. Galasso, F.S. Structure of Perovskite-Type Compounds. In Structure, Properties and Preparation of Perovskite-Type Compounds, 1st ed.; Smoluchowski, R., Kurti, N., Eds.; Elsevier Pergamon Press Inc.: Oxford, UK, 1969; pp. 2–49. [Google Scholar]
  11. Goldschmidt, V.M. Die gesetze der krystallochemie. Naturwissenschaften 1926, 14, 477–485. [Google Scholar] [CrossRef]
  12. Pena, M.A.; Fierro, J.L.G. Chemical structures and performance of perovskite oxides. Chem. Rev. 2001, 101, 1981–2017. [Google Scholar] [CrossRef]
  13. Bhalla, A.S.; Guo, R.; Roy, R. The perovskite structure—A review of its role in ceramic science and technology. Mater. Res. Innov. 2000, 4, 3–26. [Google Scholar] [CrossRef]
  14. Poglitsch, A.; Weber, D. Dynamic disorder in methylammoniumtrihalogenoplumbates (II) observed by millimeter-wave spectroscopy. J. Chem. Phys. 1987, 87, 6373–6378. [Google Scholar] [CrossRef]
  15. Li, W.; Wang, Z.; Deschler, F.; Gao, S.; Friend, R.H.; Cheetham, A.K. Chemically diverse and multifunctional hybrid organic–inorganic perovskites. Nat. Rev. Mater. 2017, 2, 1–18. [Google Scholar] [CrossRef]
  16. Weber, D. CH3NH3PbX3, a Pb(II)-system with cubic perovskite structure. J. Chem. Sci. B 1978, 33, 1443–1445. [Google Scholar] [CrossRef]
  17. Kieslich, G.; Sun, S.; Cheetham, A.K. An extended tolerance factor approach for organic–inorganic perovskites. Chem. Sci. 2015, 6, 3430–3433. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  18. Chen, J.C.; Wu, J.M. Dielectric properties and ac conductivities of dense single-phased BiFeO3 ceramics. Appl. Phys. Lett. 2007, 91, 182903. [Google Scholar] [CrossRef]
  19. Nakayam, S. LaFeO3 perovskite-type oxide prepared by oxide-mixing, co-precipitation and complex synthesis methods. J. Mater. Sci. 2001, 36, 5643–5648. [Google Scholar] [CrossRef]
  20. Muneeswaran, M.; Jegatheesan, P.; Giridharan, N.V. Synthesis of nanosized BiFeO3 powders by co-precipitation method. J. Exp. Nanosci. 2012, 8, 341–346. [Google Scholar] [CrossRef]
  21. Dhakad, M.; Rayalu, S.S.; Kumar, R.; Doggali, P.; Bakardjieva, S.; Subrt, J.; Mitsuhashi, T.; Haneda, H.; Labhsetwar, N. Low cost, ceria promoted perovskite type catalysts for diesel soot oxidation. Catal. Lett. 2008, 121, 137–143. [Google Scholar] [CrossRef]
  22. Huang, F.; Sun, X.; Zheng, Y.; Xiao, Y.; Zheng, Y. Facile coprecipitation synthesis of La0.6Sr0.4MnO3 perovskites with high surface area. Mater. Lett. 2017, 210, 287–290. [Google Scholar] [CrossRef]
  23. Gosavi, P.V.; Biniwale, R.B. Pure phase LaFeO3 perovskite with improved surface area synthesized using different routes and its characterization. Mater. Chem. Phys. 2010, 119, 324–329. [Google Scholar] [CrossRef]
  24. Delmon, B.; Devillers, M. Solid State Reactions. In Handbook of Heterogeneous Catalysis, 2nd ed.; Ertl, G., Knozinger, H., Schuth, F., Weitkamp, J., Eds.; Wiley: Weinheim, Germany, 2008; pp. 295–318. [Google Scholar]
  25. Esposito, S. “Traditional” sol-gel chemistry as a powerful tool for the preparation of supported metal and metal oxide catalysts. Materials 2019, 12, 668. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  26. Shu, J.; Kaliaguine, S. Well-dispersed perovskite-type oxidation catalysts. Appl. Catal. B 1998, 16, 303–308. [Google Scholar] [CrossRef]
  27. Rida, K.; Benabbas, A.; Bouremmad, F.; Pen, M.A.; Sastre, E.; Martınez-Arias, A. Effect of calcination temperature on the structural characteristics and catalytic activity for propene combustion of sol–gel derived lanthanum chromite perovskite. Appl. Catal. A 2007, 327, 173–179. [Google Scholar] [CrossRef]
  28. Zhong, Z.; Chen, L.; Yan, Q.; Fu, X. Study on the preparation of nanometer perovskite-type complex oxide LaFeO3 by sol-gel method. Stud. Surf. Sci. Catal. 1995, 91, 647–655. [Google Scholar] [CrossRef]
  29. Wei, J.; Xue, D. Low-temperature synthesis of BiFeO3 nanoparticles by ethylenediaminetetraacetic acid complexing sol–gel process. Mater. Res. Bull. 2008, 43, 3368–3373. [Google Scholar] [CrossRef]
  30. Ghosh, S.; Dasgupta, S.; Sen, A.; Maiti, H.S. Low temperature synthesis of bismuth ferrite nanoparticles by a ferrioxalate precursor method. Mater. Res. Bull. 2005, 40, 2073–2079. [Google Scholar] [CrossRef]
  31. Wang, X.; Zhang, Y.; Wu, Z. Magnetic and optical properties of multiferroic bismuth ferrite nanoparticles by tartaric acid-assisted sol–gel strategy. Mater. Lett. 2010, 64, 486–488. [Google Scholar] [CrossRef]
  32. Liu, T.; Xu, Y. Synthesis of nanocrystalline LaFeO3 powders via glucose sol–gel route. Mater. Chem. Phys. 2011, 129, 1047–1050. [Google Scholar] [CrossRef]
  33. Dhal, G.; Dey, S.; Mohan, D.; Prasa, R. Solution combustion synthesis of perovskite-type catalysts for diesel engine exhaust gas purification. Mater. Today Proc. 2017, 4, 10489–10493. [Google Scholar] [CrossRef]
  34. Peñalva, J.; Lazo, A. Synthesis of bismuth ferrite BiFeO3 by solution combustion method. In Journal of Physics: Conference Series, Proceedings of the XVI Meeting of Physics, Lima, Peru, 2–4 August 2017; Ochoa, R., Salinas, E., Blas, H., Eds.; IOP Publishing: Bristol, UK, 2017; pp. 195–202. [Google Scholar]
  35. Biamino, S.; Badini, C. Combustion synthesis of lanthanum chromite starting from water solutions: Investigation of process mechanism by DTA–TGA–MS. J. Eur. Ceram. Soc. 2004, 24, 3021–3034. [Google Scholar] [CrossRef]
  36. Russo, N.; Fino, D.; Saracco, G.; Specchia, V. Promotion effect of Au on perovskite catalysts for the regeneration of diesel particulate filters. Catal. Today. 2008, 137, 306–311. [Google Scholar] [CrossRef]
  37. Shandilya, M.; Rai, R.; Singh, J. Review: Hydrothermal technology for smart materials. Appl. Ceram. 2016, 115, 354–376. [Google Scholar] [CrossRef]
  38. Niu, F.; Gao, T.; Zhang, N.; Chen, Z.; Huang, Q.; Qin, L.; Sun, X.; Huang, Y. Hydrothermal synthesis of BiFeO3 nanoparticles for visible light photocatalytic applications. Mater. Res. Bull. 2014, 59, 6–12. [Google Scholar] [CrossRef]
  39. Zheng, W.; Liu, R.; Peng, D.; Meng, G. Hydrothermal synthesis of LaFeO3 under carbonate-containing medium. Mater. Lett. 2000, 43, 19–22. [Google Scholar] [CrossRef]
  40. Assirey, E.A.R. Perovskite synthesis, properties and their related biochemical and industrial application. Saudi Pharm. J. 2019, 27, 817–829. [Google Scholar] [CrossRef]
  41. Perejón, A.; Murafa, N.; Sánchez-Jiménez, P.E.; Criado, J.M.; Subrt, J.; Diánez, M.J.; Pérez-Maqued, L.A. Direct mechanosynthesis of pure BiFeO3 perovskite nanoparticles: Reaction mechanism. J. Mater. Chem. C 2013, 1, 3551–3562. [Google Scholar] [CrossRef] [Green Version]
  42. Sorescu, M.; Xu, T.; Burnett, J.D.; Aitken, J.A. Investigation of LaFeO3 perovskite growth mechanism through mechanical ball milling of lanthanum and iron oxides. J. Mater. Sci. 2011, 46, 6709–6717. [Google Scholar] [CrossRef]
  43. Amador, C.; Martin de Juan, L. Strategies for structured particulate systems design. Comput. Aided Chem. Eng. 2016, 39, 509–579. [Google Scholar] [CrossRef]
  44. Popescu, I.; Sandulescu, I.; Redey, A.; Marcu, I.C. Study of the catalytic activity–semiconductive properties relationship for BaTiO3 and PbTiO3 perovskites, catalysts for methane combustion. Catal. Lett. 2011, 141, 445–451. [Google Scholar] [CrossRef]
  45. Arai, H.; Yamada, T.; Eguchi, K.; Seiyama, T. Catalytic combustion of methane over various perovskite-type oxides. Appl. Catal. 1986, 26, 265–276. [Google Scholar] [CrossRef]
  46. Majid, A.; Tunney, J.; Argue, S.; Post, M. The effect of preparation method and calcination temperature on the crystallite size and surface area of perovskite-type SrFeOx. J. Sol Gel Sci. Technol. 2004, 32, 323–326. [Google Scholar] [CrossRef] [Green Version]
  47. Saracco, G.; Scibilia, G.; Iannibello, A.; Baldi, G. Methane combustion on Mg-doped LaCrO, perovskite catalysts. Appl. Catal. B 1996, 8, 229–244. [Google Scholar] [CrossRef]
  48. Rosso, I.; Saracco, G.; Specchia, V.; Garrone, E. Sulphur poisoning of LaCr0.5−xMnxMg0.5O3 ∙ yMgO catalysts for methane combustion. Appl. Catal. B 2003, 40, 195–205. [Google Scholar] [CrossRef]
  49. Salomonsson, P.; Griffin, T.; Kasemo, B. Oxygen desorption and oxidation-reduction kinetics with methane and carbon monoxide over perovskite type metal oxide catalysts. Appl. Catal. A 1993, 194, 175–197. [Google Scholar] [CrossRef]
  50. Lisi, L.; Bagnasco, G.; Ciambelli, P.; De Rossi, S.; Porta, P.; Russo, G.; Turco, M. Perovskite-type oxides, II. Redox properties of LaMn1-xCuxO3 and LaCo1-xCuxO3 and methane catalytic combustion. J. Solid State Chem. 1999, 146, 176–183. [Google Scholar] [CrossRef]
  51. Ciambellia, P.; Palma, V.; Tikhov, S.F.; Sadykov, V.A.; Isupova, L.A.; Lisi, L. Catalytic activity of powder and monolith perovskites in methane combustion. Catal. Today 1999, 47, 199–207. [Google Scholar] [CrossRef]
  52. Alifanti, M.; Kirchnerova, J.; Delmon, B.; Klvana, D. Methane and propane combustion over lanthanum transition-metal perovskites: Role of oxygen mobility. Appl. Catal. A 2004, 262, 167–176. [Google Scholar] [CrossRef]
  53. Kirchnerova, J.; Klvana, D.; Vaillancourt, J.; Chaouki, J. Evaluation of some cobalt and nickel based perovskites prepared by freeze-drying as combustion catalysts. Catal. Lett. 1993, 21, 77–87. [Google Scholar] [CrossRef]
  54. Kirchenerov, J.; Klvana, D. Preparation and characterization of high surface perovskite electrocatalysts. Int. J. Hydrogen Energy 1994, 19, 501–506. [Google Scholar] [CrossRef]
  55. Cimino, S.; Lisi, S.; De Rossi, S.; Faticanti, M.; Porta, P. Methane combustion and CO oxidation on LaAl1−xMnxO3 perovskite-type oxide solid solutions. Appl. Catal. B 2003, 43, 397–406. [Google Scholar] [CrossRef]
  56. Baiker, A.; Marti, P.E.; Keusch, P.; Fritsch, E.; Relier, A. Influence of the A-Site cation in ACoO3 (A = La, Pr, Nd, perovskite-type oxides on catalytic activity for methane combustion. J. Catal. 1994, 146, 268–276. [Google Scholar] [CrossRef]
  57. Busca, G.; Daturi, M.; Finocchio, E.; Lorenzelli, V.; Ramis, G.; Willey, R.J. Transition metal mixed oxides as combustion catalysts: Preparation, characterization and activity mechanisms. Catal. Today 1997, 33, 239–249. [Google Scholar] [CrossRef]
  58. Kirchnerova, J.; Alifanti, M.; Delmon, B. Evidence of phase cooperation in the LaCoO3–CeO2–Co3O4 catalytic system in relation to activity in methane combustion. Appl. Catal. A 2002, 231, 65–80. [Google Scholar] [CrossRef]
  59. Yi, N.; Cao, Y.; Su, Y.; Dai, W.L.; He, H.Y.; Fan, K.N. Nanocrystalline LaCoO3 perovskite particles confined in SBA-15 silica as a new efficient catalyst for hydrocarbon oxidation. J. Catal. 2005, 230, 249–253. [Google Scholar] [CrossRef]
  60. Yamazoe, N.; Teraoka, Y. Oxidation catalysis of perovskites—Relationships to bulk structure and composition (valency, defect, etc.). Catal. Today 1990, 8, 175–199. [Google Scholar] [CrossRef]
  61. Rossetti, I.; Biffi, C.; Forni, L. Oxygen non-stoichiometry in perovskitic catalysts: Impact on activity for the flameless combustion of methane. Chem. Eng. J. 2010, 162, 768–775. [Google Scholar] [CrossRef] [Green Version]
  62. Koirala, R.; Pratsinis, S.E.; Baiker, A. Synthesis of catalytic materials in flames: Opportunities and challenges. Chem. Soc. Rev. 2016, 45, 3053–3068. [Google Scholar] [CrossRef] [Green Version]
  63. Milt, V.G.; Spretz, R.; Ulla, M.A.; Lambardo, E.A. The nature of active sites for the oxidation of methane on La-based perovskites. Catal. Lett. 1996, 42, 57–63. [Google Scholar] [CrossRef]
  64. Wei, X.; Hug, P.; Figi, R.; Trottmann, M.; Weidenkaff, A.; Ferri, D. Catalytic combustion of methane on nano-structured perovskite-type oxides fabricated by ultrasonic spray combustion. Appl. Catal. B 2010, 94, 27–37. [Google Scholar] [CrossRef]
  65. Rossetti, I.; Forni, L. Catalytic flameless combustion of methane over perovskites prepared by flame–hydrolysis. Appl. Catal. B 2001, 33, 345–352. [Google Scholar] [CrossRef]
  66. Ding, Y.; Wang, S.; Zhang, L.; Chen, Z.; Wang, M.; Wang, S. A facile method to promote LaMnO3 perovskite catalyst for combustion of methane. Catal. Commun. 2017, 97, 88–92. [Google Scholar] [CrossRef]
  67. Lu, Y.; Eyssler, A.; Otal, E.H.; Matam, S.K.; Brunko, O.; Weidenkaff, A.; Ferri, D. Influence of the synthesis method on the structure of Pd-substituted perovskite catalysts for methane oxidation. Catal. Today 2013, 208, 42–47. [Google Scholar] [CrossRef]
  68. Lee, N.H.; Kim, J.H.; Oh, H.J.; Kim, S.J. In-Situ preparation of highly crystallized ZnO manopowders by ultrasonic spraying combustion method. Solid State Phenom. 2007, 121–123, 1269–1272. [Google Scholar] [CrossRef]
  69. Marti, P.E.; Baiker, A. Influence of the A-site cation in AMnO3+x and AFeO3+x (A = La, Pr, Nd and Gd) perovskite-type oxides on the catalytic activity for methane combustion. Catal. Lett. 1994, 26, 71–84. [Google Scholar] [CrossRef]
  70. Cimino, S.; Lisi, L.; Pirone, R.; Russo, G.; Turco, M. Methane combustion on perovskites-based structured catalysts. Catal. Today 2000, 59, 19–31. [Google Scholar] [CrossRef]
  71. Marchetti, L.; Forni, L. Catalytic combustion of methane over perovskites. Appl. Catal. B 1998, 15, 179–187. [Google Scholar] [CrossRef]
  72. Ciambelli, P.; Cimino, S.; De Rossi, S.; Faticanti, M.; Lisi, L.; Minelli, G.; Pettiti, I.; Porta, P.; Russo, G.; Turco, M. AMnO3 (A = La, Nd, Sm) and Sm1−xSrxMnO3 perovskites as combustion catalysts: Structural, redox and catalytic properties. Appl. Catal. B 2000, 24, 243–253. [Google Scholar] [CrossRef]
  73. Marti, P.E.; Maciejewski, M.; Baiker, A. Methane combustion over La0.8Sr0.2MnO3+x supported on MAl2O4 (M = Mg, Ni and Co) spinels. Appl. Catal. B 1994, 4, 225–235. [Google Scholar] [CrossRef]
  74. Wang, Y.; Arandiyan, H.; Tahini, H.A.; Scott, J.; Tan, X.; Dai, H.; Gale, J.D.; Rohl, A.L.; Smith, S.C.; Amal, R. The controlled disassembly of mesostructured perovskites as an avenue to fabricating high performance nanohybrid catalysts. Nat. Commun. 2017, 8, 15553. [Google Scholar] [CrossRef]
  75. Daturi, M.; Busca, G.; Groppi, G.; Forzatti, P. Preparation and characterisation of SrTil-x-yZrxMnyO3 solid solution powders in relation to their use in combustion catalysis. Appl. Catal. B 1997, 12, 325–337. [Google Scholar] [CrossRef]
  76. Pecchi, G.; Reyes, P.; Zamora, R.; Campos, C.; Cadus, L.E.; Barbero, B.P. Effect of the preparation method on the catalytic activity of La1-xCaxFeO3 perovskite-type oxides. Catal. Today 2008, 133–135, 420–427. [Google Scholar] [CrossRef]
  77. Ciambelli, P.; Cimino, S.; De Rossi, S.; Lisi, L.; Minelli, G.; Porta, P.; Russo, G. AFeO3 (A = La, Nd, Sm) and LaFe1−xMgxO3 perovskites as methane combustion and CO oxidation catalysts: Structural, redox and catalytic properties. Appl. Catal. B 2001, 29, 239–250. [Google Scholar] [CrossRef]
  78. Ciambelli, P.; Cimino, S.; Lisi, L.; Faticanti, M.; Minelli, G.; Pettiti, I.; Porta, P. La, Ca and Fe oxide perovskites: Preparation, characterization and catalytic properties for methane combustion. Appl. Catal. B 2001, 33, 193–203. [Google Scholar] [CrossRef]
  79. Pecchi, G.; Jiliberto, M.G.; Buljan, A.; Delgado, E.J. Relation between defects and catalytic activity of calcium doped LaFeO3 perovskite. Solid State Ion. 2011, 187, 27–32. [Google Scholar] [CrossRef]
  80. Kucharczyk, B.; Tylus, W. Partial substitution of lanthanum with silver in the LaMnO3 perovskite: Effect of the modification on the activity of monolithic catalysts in the reactions of methane and carbon oxide oxidation. Appl. Catal. A 2008, 335, 28–36. [Google Scholar] [CrossRef]
  81. Kucharczyk, B.; Okal, J.; Tylus, W.; Winiarski, J.; Szczygieł, B. The effect of the calcination temperature of LaFeO3 precursors on the properties and catalytic activity of perovskite in methane oxidation. Ceram. Int. 2019, 45, 2779–2788. [Google Scholar] [CrossRef]
  82. Specchia, S.; Toniato, G. Natural gas combustion catalysts for environmental-friendly domestic burners. Catal. Today 2009, 147, 99–106. [Google Scholar] [CrossRef]
  83. Klvanaa, D.; Kirchnerova, J.; Chaouki, J.; Delval, J.; Yaici, W. Fiber-supported perovskites for catalytic combustion of natural gas. Catal. Today 1999, 47, 115–121. [Google Scholar] [CrossRef]
  84. Pattini, F. Growth of Oxide Thin Films for Energy Devices by Pulsed Electron Deposition. Ph.D. Thesis, Università degli Studi di Parma, Parma, Italy, 2009. Available online: https://www.repository.unipr.it/handle/1889/1001 (accessed on 8 November 2020).
  85. Norton, D.P. Pulsed Laser Deposition of Complex Materials: Progress Toward Applications. In Pulsed Laser Deposition of Thin films; Eason, R., Ed.; Wiley: Hoboken, NJ, USA, 2007; pp. 3–28. [Google Scholar]
  86. Constantinescu, C.; Dinescu, M. Thin films development by pulsed laser-assisted deposition. Physics AUC 2010, 20, 43–56. [Google Scholar]
  87. Marozau, I.; Shkabko, A.; Dinescu, G.; Döbeli, M.; Lippert, T.; Logvinovich, D.; Malllepell, M.; Weidenkaff, A.; Wokaun, A. RF-plasma assisted pulsed laser deposition of nitrogen-doped SrTiO3 thin films. Appl. Phys. A 2008, 93, 721–727. [Google Scholar] [CrossRef]
  88. Beckers, L.; Schubert, J.; Zander, W.; Ziesmann, J.; Eckau, A.; Leinenbach, P.; Buchal, C. Structural and optical characterization of epitaxial wave guiding BaTiO3 thin films on MgO. J. Appl. Phys. 1998, 83, 3305–3310. [Google Scholar] [CrossRef] [Green Version]
  89. Mi, S.B.; Jia, C.L.; Heeg, T.; Trithaveesak, O.; Schubert, J.; Urban, K. Heterostructures of BaTiO3 bilayer films grown on SrTiO3 (001) under different oxygen pressures. J. Cryst. Growth 2005, 283, 425–430. [Google Scholar] [CrossRef]
  90. Berini, B.; Evain, M.; Fouchet, A.; Dumont, Y.; Popova, E.; Keller, N. In situ optical characterization of metal-insulator transition in LaNiO3 and SrTiO3 perovskites in pulsed laser deposition chamber. Phase Transit. 2011, 84, 501–508. [Google Scholar] [CrossRef]
  91. Cillessen, J.F.M.; Wolf, R.M.; Giesbers, J.B.; Blom, P.W.M.; Grosse-Holz, K.O.; Pastoor, E. Growth, structuring and characterisation of all-oxide thin film devices prepared by pulsed laser deposition. Appl. Surf. Sci. 1996, 96–98, 744–751. [Google Scholar] [CrossRef] [Green Version]
  92. Francis, A.; Salvador, P. Synthesis, structures, and physical properties of yttrium-doped strontium manganese oxide films. Mat. Res. Soc. Symp. Proc. 2002, 718, D9.4.1–D9.4.6. [Google Scholar] [CrossRef] [Green Version]
  93. Hawley, M.E.; Adams, C.D.; Arendt, P.N.; Brosha, E.L.; Garzon, F.H.; Houlton, R.J.; Hundley, M.F.; Heffner, R.H.; Jia, Q.X.; Neumeier, J.; et al. CMR films structure as a function of growth and processing. J. Cryst. Growth 1997, 174, 455–463. [Google Scholar] [CrossRef]
  94. Surthi, S.; Kotru, S.; Pandey, R.K. Characterization of integrated heterostructures of SbSI–La0.67Ca0.33MnO3. Mater. Lett. 2003, 57, 3455–3460. [Google Scholar] [CrossRef]
  95. Chang, J.; Lee, J.W.; Kim, S.K. Layer-by-layer growth of SrFeO3-δ thin films on atomically flat single-terminated SrRuO3/SrTiO3 (111) surfaces. J. Cryst. Growth 2010, 312, 621–623. [Google Scholar] [CrossRef]
  96. Sahner, K.; Moos, R.; Matam, M.; Tunney, J.J.; Post, M. Hydrocarbon sensing with thick and thin film p-type conducting perovskite materials. Sens. Actuators B Chem. 2005, 108, 102–112. [Google Scholar] [CrossRef] [Green Version]
  97. Pignolet, A.; Satyalakshmi, K.M.; Alexe, M.; Zakharov, N.D.; Harnagea, C.; Senz, S.; Hesse, D.; Gösele, U. Epitaxial bismuth-layer-structured perovskite ferroelectric thin films grown by pulsed laser deposition. Integr. Ferroelectr. 1999, 26, 21–29. [Google Scholar] [CrossRef]
  98. Coy, L.E.; Rebled, J.; Ventura, J.; Yate, L.; Ferrater, C.; Langenberg, E.; Polo, M.C.; Xuriguera, E.; Peiro, F.; Varela, M. Crystalline domains in epitaxial Y(Ni0.5Mn0.5)O3 thin films grown by PLD on different STO substrates. Appl. Surf. Sci. 2015, 324, 114–122. [Google Scholar] [CrossRef]
  99. Hassini, A.; Gervais, M.; Coulon, J.; Phuoc, V.T.; Gervais, F. Synthesis of Ca0.25Cu0.75TiO3 and infrared characterization of role played by copper. Mater. Sci. Eng. B 2001, 87, 164–168. [Google Scholar] [CrossRef]
  100. Khan, M.A.; Comyn, T.P.; Bell, A.J. Deposition of PbTiO3 films on Pt/Si substrates using pulsed laser deposition. J. Eur. Ceram. Soc. 2008, 28, 591–597. [Google Scholar] [CrossRef]
  101. Wang, Z.J.; Kokawa, H.; Maeda, R. Growth of lead zirconate titanate thin films by hybrid processing: Sol–gel method and pulsed-laser deposition. J. Cryst. Growth 2004, 262, 359–365. [Google Scholar] [CrossRef]
  102. Yun, K.Y.; Ricinschi, D.; Kanashima, T.; Noda, M.; Okuyama, M. Giant ferroelectric polarization beyond 150 C/cm2 in BiFeO3 thin film. Jpn. J. Appl. Phys. 2004, 43, 647–648. [Google Scholar] [CrossRef]
  103. Béa, H.; Bibes, M.; Barthélémy, A.; Bouzehouane, K.; Jacquet, E.; Khodan, A.; Contour, J.P.; Fusil, S.; Wyczisk, F.; Forget, A.; et al. Influence of parasitic phases on the properties of BiFeO3 epitaxial thin films. Appl. Phys. Lett. 2005, 87, 072508. [Google Scholar] [CrossRef] [Green Version]
  104. Zhang, G.; Cheng, J.; Chen, R.; Yu, S.; Meng, Z. Preparation of BiFeO3 thin films by pulsed laser deposition method. Trans. Nonferrous. Met. Soc. China 2006, 16, 123–125. [Google Scholar] [CrossRef]
  105. Harnagea, C.; Cojocaru, C.V.; Gautreau, O.; Nechache, R.; Normandin, F.; Veres, T.; Pignolet, A. Piezoresponse force microscopy of PLD-grown multiferroic BiFeO3 films and mesostructures. Integr. Ferroelectr. 2006, 83, 1–12. [Google Scholar] [CrossRef]
  106. Go, H.; Wakiya, N.; Funakubo, H.; Satoh, K.; Kondo, M.; Cross, J.S.; Maruyama, K.; Mizutani, N.; Shinozaki, K. Effect of oxygen annealing on ferroelectricity of BiFeO3 thin films formed by Pulsed Laser Deposition. Jpn. J. Appl. Phys. 2007, 46, 3491–3494. [Google Scholar] [CrossRef]
  107. Himcinschi, C.; Vrejoiu, I.; Friedrich, M.; Ding, L.; Cobet, C.; Esser, N.; Alexe, M.; Zahn, D.R.T. Optical characterisation of BiFeO3 epitaxial thin films grown by pulsed-laser deposition. Phys. Status Solidi C 2010, 7, 296–299. [Google Scholar] [CrossRef]
  108. Vrejoiu, I.; Alexe, M.; Hesse, D.; Gösele, U. Functional perovskites—From epitaxial films to nanostructured arrays. Adv. Funct. Mater. 2008, 18, 3892–3906. [Google Scholar] [CrossRef]
  109. Ravalia, A.; Vagadia, M.; Trivedi, P.; Solanki, P.S.; Asokan, K.; Ojha, S.; Thakur, O.P.; Choudhary, R.J.; Phase, D.M.; Kuberkar, D.G. Role of oxygen in multiferroic behavior of BiFeO3 films grown on 0.2% Nb doped SrTiO3. Solid State Commun. 2013, 169, 10–13. [Google Scholar] [CrossRef]
  110. Prashanthi, K.; Gupta, M.; Tsui, Y.Y.; Thundat, T. Effect of annealing atmosphere on microstructural and photoluminescence characteristics of multiferroic BiFeO3 thin films prepared by pulsed laser deposition technique. Appl. Phys. A 2012, 110, 903–907. [Google Scholar] [CrossRef]
  111. Zhao, L.; Lu, Z.; Zhang, F.; Tian, G.; Song, X.; Li, Z.; Huang, K.; Zhang, Z.; Qin, M.; Wu, S.; et al. Current rectifying and resistive switching in high density BiFeO3 nanocapacitor arrays on Nb-SrTiO3 substrates. Sci. Rep. 2015, 5, 9680. [Google Scholar] [CrossRef] [Green Version]
  112. Jiang, Z.Z.; Guan, Z.; Yang, N.; Xiang, P.H.; Qi, R.J.; Huang, R.; Yang, P.X.; Zhong, N.; Duan, C.G. Epitaxial growth of BiFeO3 films on SrRuO3/SrTiO3. Mater. Charact. 2017, 131, 217–223. [Google Scholar] [CrossRef]
  113. Ravaliya, K.; Ravalia, A.; Pandya, D.D.; Solanki, P.S.; Shah, N.A. Strain and morphology control over electrical behavior of pulsed laser deposited BiFeO3 films. Thin Solid Films 2018, 645, 436–443. [Google Scholar] [CrossRef]
  114. Singh, G.; Yadav, R.P.; Bhasker, H.P.; Kumar, M.; Rajput, P.; Rao, P.N.; Rai, S.K.; Singh, M.K. Investigation of fractal feature of multiferroic BiFeO3 thin films deposited on different substrates. Mater. Res. Express 2018, 5, 126405. [Google Scholar] [CrossRef]
  115. Lee, D.; Kim, M.G.; Ryu, S.; Jang, H.M.; Lee, S.G. Epitaxially grown La-modified BiFeO3 magnetoferroelectric thin films. Appl. Phys. Lett. 2005, 86, 222903. [Google Scholar] [CrossRef] [Green Version]
  116. Sahoo, K.K.; Katoch, R.; Brajesh, K.; Garg, A.; Gupta, R. Improved ferroelectric response of pulsed laser deposited BiFeO3-PbTiO3 thin films around morphotropic phase boundary with interfacial PbTiO3 buffer layer. J. Appl. Phys. 2020, 127, 064101. [Google Scholar] [CrossRef]
  117. Khan, M.A.; Comyn, T.P.; Bell, A.J. Growth and characterization of tetragonal bismuth ferrite–lead titanate thin films. Acta Mater. 2008, 56, 2110–2118. [Google Scholar] [CrossRef]
  118. Talley, K.R.; Barron, S.C.; Nguyen, N.; Wong-Ng, W.; Martin, J.; Zhang, Y.L.; Song, X. Thermoelectric properties of the LaCoO3-LaCrO3 system using a high-throughput combinatorial approach. Solid State Sci. 2017, 64, 7–12. [Google Scholar] [CrossRef] [Green Version]
  119. Choi, W.S.; Marton, Z.; Jang, S.Y.; Moon, S.J.; Jeon, B.C.; Shin, J.H.; Seo, S.S.A.; Noh, T.W.; Myung-Whun, K.; Lee, H.N.; et al. Effects of oxygen-reducing atmosphere annealing on LaMnO3 epitaxial thin films. J. Phys. D Appl. Phys. 2009, 42, 165401. [Google Scholar] [CrossRef]
  120. Zhang, X.; Zhang, A.M.; Xie, W.M.; Lin, J.G.; Wu, X.S. Effect of strain-modulated lattice distortion on the magnetic properties of LaMnO3 films. Physica B Condens. Matter. 2015, 476, 114–117. [Google Scholar] [CrossRef]
  121. Debnath, J.C.; Strydom, A.M. Large low field magneto-resistance and temperature coefficient of resistance in La0.8Ca0.2MnO3 epitaxial thin film. J. Alloys Compd. 2015, 621, 7–11. [Google Scholar] [CrossRef]
  122. Xiong, H.F.; Cheng, T.D.; Tang, X.G.; Chen, J.; Liu, Q.X. X-ray photoelectron spectroscopy of (La0.7Sr0.3)MnO3 thin films prepared by Pulsed Laser Deposition. Adv. Mat. Res. 2011, 284–286, 2191–2197. [Google Scholar] [CrossRef]
  123. Brown, G.W.; Jia, Q.X.; Peterson, E.J.; Hristova, D.K.; Hundley, M.F.; Thompson, J.D.; Maggiore, C.J.; Tesmer, J.; Hawley, M.E. Growth and magnetic structure of La0.67Sr0.33MnO3 films. Mat. Res. Soc. Symp. Proc. 1997, 474, 179–184. [Google Scholar] [CrossRef]
  124. Iwata, N.; Huijben, M.; Rijnders, G.; Yamamoto, H.; Blank, D.H.A. Growth of CaFeOX/LaFeO3 Superlattice on SrTiO3(100) Substrates. Mat. Res. Soc. Symp. Proc. 2011, 1292, 125–130. [Google Scholar] [CrossRef]
  125. Park, J.; Kim, Y.; Lee, D.; Song, J.H.; Park, J.H. Twin-free epitaxial LaFeO3 films grown on orthorhombic GdScO3(110) substrates. J. Korean Phys. Soc. 2020, 76, 273–276. [Google Scholar] [CrossRef]
  126. Tsuchiya, Y.; Norota, K.; Watabe, Y.; Kuroda, T.; Iwata, N.; Hashimoto, T.; Yamamoto, H. Growth difference of LaFeO3 thin films by pulsed laser deposition method using the targets prepared by pechini and conventional solid solution methods. Trans. Mat. Res. Soc. Jap. 2012, 37, 369–372. [Google Scholar] [CrossRef] [Green Version]
  127. Liu, H.; Zhu, J.; Xiao, D. Preparation and characterization of LaFeO3 thin films on (100) SrTiO3 substrates by pulsed laser deposition. J. Adv. Dielectr. 2011, 1, 363–367. [Google Scholar] [CrossRef]
  128. Andrei, F.; Boerasu, I.; Moldovan, A.; Dinescu, M.; Ion, V.; Mihailescu, C.; Scarisoreanu, N.D.; Leca, V. The effects of the oxygen content on the photoelectrochemical properties of LaFeO3 perovskite thin films obtained by pulsed laser deposition. Appl. Phys. A 2019, 125, 807. [Google Scholar] [CrossRef]
  129. Manzoor, S.; Somvanshi, A.; Husain, S. Structural analysis of LaFeO3 thin films grown on SrTiO3 and LaAlO3 substrates. AIP Conf. Proc. 2018, 1942, 80011. [Google Scholar] [CrossRef]
  130. Cieniek, Ł.; Kopia, A.; Kusiński, J.; Kowalski, K.; Moskalewicz, T.; Cyza, A.; Maziarz, W. Microstructure and selected mechanical and electrical property analysis of Sr-doped LaCoO3 perovskite thin films deposited by the PLD technique. Int. J. Mater. Res. 2018, 110, 32–41. [Google Scholar] [CrossRef]
  131. Kopia, A.; Cieniek, Ł.; Kowalski, K.; Kusiński, J. Influence of Sr-Doping on the structure of LaCoO3 thin films prepared by pulsed laser deposition. Solid State Phenom. 2015, 231, 19–24. [Google Scholar] [CrossRef]
  132. Cyza, A.; Cieniek, Ł.; Kopia, A. Perovskite La1-xSrxFeO3 thin films deposited by laser ablation process. Arch. Metall. Mater. 2016, 61, 1063–1067. [Google Scholar] [CrossRef] [Green Version]
  133. Dash, U.; Acharya, S.K.; Cho, S.W.; Lee, S.; Lee, K.; Chae, S.C.; Cho, M.R.; Jung, C.U. Large magnetoresistance in LaFeO3 substituted SrRuO3 epitaxial thin films. J. Alloys Compd. 2017, 724, 549–554. [Google Scholar] [CrossRef]
  134. Kolte, J.; Daryapurkar, A.S.; Agarwal, M.; Gulwade, D.D.; Gopalan, P. Effect of substrate temperature on the structural and electrical properties of La and Mn co-doped BiFeO3 thin films. Thin Solid Films 2016, 619, 308–316. [Google Scholar] [CrossRef]
  135. Izumi, H.; Yoshimura, T.; Fujimura, N. Growth and ferroelectric properties of La and Al codoped BiFeO3 epitaxial films. J. Appl. Phys. 2017, 121, 174102. [Google Scholar] [CrossRef]
  136. Zhao, Y.; Srivastava, R.; Fournier, P.; Smolyaninova, V.; Rajeswari, M.; Wu, T.; Li, Z.Y.; Greene, R.L.; Venkatesan, T. Transport and magnetic properties of La0.8Ce0.2MnO3 thin films grown by pulsed laser deposition. J. Magn. Magn. Mater. 2000, 220, 161–166. [Google Scholar] [CrossRef]
  137. Lin, T.K.; Chang, H.W.; Chen, B.A.; Wei, D.H.; Wang, C.R.; Tu, C.S. Effect of Pr substitution on the structure, nanomechanical and multiferroic characterizations of Bi1-xPrxFeO3 polycrystalline films. Surf. Coat. Technol. 2020, 393, 125728. [Google Scholar] [CrossRef]
  138. Bowman, R.M.; Catalan, G.; Corbett, M.H.; O’Neill, D.; Gregg, J.M. PLD of metal insulator and relaxor electroceramic thin films. In 11th International School on Quantum Electronics: Laser Physics and Applications, Proceedings of the SPIE 2001; Atanasov, P.A., Cartaleva, C., Eds.; SPIE: Bellingham, WA, USA, 2001; Volume 4397, pp. 273–284. [Google Scholar] [CrossRef]
  139. Shiyani, T.; Shekhada, K.G.; Savaliya, C.R.; Markna, J.H. Tuning the metal-insulator transition in NdNiO3 thin films. AIP Conf. Proc. 2017, 1837, 040006. [Google Scholar] [CrossRef]
  140. Puli, V.S.; Pradhan, D.K.; Katiyar, R.K.; Coondoo, I.; Panwar, N.; Misra, P.; Chrisey, D.B.; Scott, J.F.; Katiyar, R.S. Photovoltaic effect in transition metal modified polycrystalline BiFeO3 thin films. J. Phys. D Appl. Phys. 2014, 47, 075502. [Google Scholar] [CrossRef] [Green Version]
  141. Lin, T.K.; Shen, C.Y.; Kao, C.C.; Chang, C.F.; Chang, H.W.; Wang, C.R.; Tu, C.S. Structural evolution, ferroelectric, and nanomechanical properties of Bi1-xSmxFeO3 films (x = 0.05–0.16) on glass substrates. J. Alloys Compd. 2019, 787, 397–406. [Google Scholar] [CrossRef]
  142. Yu, Y.; Zhang, X.; Yang, J.J.; Wang, J.W.; Zhao, Y.G. Microstructure and strain relaxation of orthorhombic TmMnO3 epitaxial thin films. J. Cryst. Growth 2012, 338, 280–282. [Google Scholar] [CrossRef]
  143. Caricato, A.P.; Luches, A. Applications of the matrix-assisted pulsed laser evaporation method for the deposition of organic, biological and nanoparticle thin films: A review. Appl. Phys. A 2011, 105, 565–582. [Google Scholar] [CrossRef]
  144. Papavlu, A.P.; Dinca, V.; Filipescu, M.; Dinescu, M. Matrix-assisted pulsed laser evaporation of organic thin gilms: Applications in biology and chemical sensors. In Laser Ablation—From Fundamentals to Applications; Itina, T.E., Ed.; InTech: Rijeka, Croatia, 2017; pp. 171–189. [Google Scholar] [CrossRef] [Green Version]
  145. Chirisey, D.B.; Pique, A.; McGill, R.A.; Horwitz, J.S.; Ringeisen, B.R.; Bubb, D.M.; Wu, P.K. Laser deposition of polymer and biomaterial films. Chem. Rev. 2003, 103, 553–576. [Google Scholar] [CrossRef]
  146. Pique, A.; McGill, R.A.; Chrisey, D.B.; Leonhardt, D.; Mslna, T.E.; Spargo, B.J.; Callahan, J.H.; Vachet, R.W.; Chung, R.; Bucaro, M.A. Growth of organic thin films by the matrix assisted pulsed laser evaporation (MAPLE) technique. Thin Solid Films 1999, 355/356, 536–541. [Google Scholar] [CrossRef]
  147. Lippert, T.; Chrisey, D.B.; Purice, A.; Constantinescu, C.; Filipescu, M.; Scarisoreanu, N.D.; Dinescu, M. Laser processing of soft materials. Rom. Rep. Phys. 2007, 59, 483–498. [Google Scholar]
  148. Queraltó, A.; György, E.; Ivan, R.; Pérez del Pino, Á.; Frohnhoven, R.; Mathur, S. Enhanced UV-Vis photodegradation of nanocomposite reduced graphene oxide/ferrite nanofiber films prepared by laser-assisted evaporation. Crystals 2020, 10, 271. [Google Scholar] [CrossRef] [Green Version]
  149. Barraza, E.T.; Dunlap-Shohl, W.A.; Mitzi, D.B.; Stiff-Roberts, A.D. Deposition of methylammonium lead triiodide by resonant infrared matrix-assisted pulsed laser evaporation. J. Electron. Mater. 2017, 47, 917–926. [Google Scholar] [CrossRef]
  150. Dunlap-Shohl, W.A.; Barraza, E.T.; Barrette, A.; Gundogdu, K.; Stiff-Roberts, A.D.; Mitzi, D.B. MAPbI3 solar cells with absorber deposited by resonant infrared matrix-assisted pulsed laser evaporation. ACS Energy Lett. 2018, 3, 270–275. [Google Scholar] [CrossRef]
  151. Manoharan, Y.; Hosseini, S.E.; Butler, B.; Alzhahrani, H.; Senior, B.T.F.; Ashuri, T.; Krohn, J. Hydrogen fuel cell vehicles; Current status and future prospect. Appl. Sci. 2019, 9, 2296. [Google Scholar] [CrossRef] [Green Version]
  152. Fan, Z.; Sunc, K.; Wang, J. Perovskites for photovoltaics: A combined review of organic–inorganic halide perovskites and ferroelectric oxide perovskites. J. Mater. Chem. A 2015, 3, 18809–18828. [Google Scholar] [CrossRef]
  153. Lede, J.; Lapicque, F.; Villermaux, J. Production of hydrogen by direct thermal decomposition of water. Int. J. Hydrogen Energy 1983, 8, 675–679. [Google Scholar] [CrossRef]
  154. Luciani, G.; Landi, G.; Aronnea, A.; Di Benedetto, A. Partial substitution of B cation in La0.6Sr0.4MnO3 perovskites: A promising strategy to improve the redox properties useful for solar thermochemical water and carbon dioxide splitting. Solar Energy 2018, 171, 1–7. [Google Scholar] [CrossRef]
  155. Bhosale, R.R. Solar hydrogen production via thermochemical magnesium oxide–Magnesium sulfate water splitting cycle. Fuel 2020, 275, 117892. [Google Scholar] [CrossRef]
  156. Bhandari, R.; Trudewind, A.C.; Zapp, P. Life cycle assessment of hydrogen production via electrolysis–A review. J. Clean. Prod. 2014, 85, 151–163. [Google Scholar] [CrossRef]
  157. Hisatomi, T.; Domen, K. Reaction systems for solar hydrogen production via water splitting with particulate semiconductor photocatalysts. Nat. Catal. 2019, 2, 387–399. [Google Scholar] [CrossRef]
  158. Samsudin, F.R.M.; Bashiri, R.; Mohamed, M.M.; Ng, H.Y.; Sufian, S. Tailoring the morphological structure of BiVO4 photocatalyst for enhanced photoelectrochemical solar hydrogen production from natural lake water. Appl. Surf. Sci. 2020, 504, 144417. [Google Scholar] [CrossRef]
  159. Mueller-Langer, F.; Tzimas, E.; Kaltschmitt, M.; Peteves, S. Techno-economic assessment of hydrogen production processes for the hydrogen economy for the short and medium term. Int. J. Hydrogen Energy 2007, 32, 3797–3810. [Google Scholar] [CrossRef]
  160. Ramachandran, R.; Menon, R.K. An overview of industrial uses of hydrogen. Int. J. Hydrogen Energy 1998, 23, 593–598. [Google Scholar] [CrossRef]
  161. Eliezer, D.; Eliaz, N.; Senkov, O.N.; Froes, F.H. Positive effects of hydrogen in metals. Mater. Sci. Eng. 2000, 280, 220–224. [Google Scholar] [CrossRef]
  162. Edwards, P.P.; Kuznetsov, V.L.; David, W.I.F.; Brandon, N.P. Hydrogen and fuel cells: Towards a sustainable energy future. Energy Policy 2008, 36, 4356–4362. [Google Scholar] [CrossRef]
  163. Abe, R. Recent progress on photocatalytic and photoelectrochemical water splitting under visible light irradiation. J. Photochem. Photobiol. C 2010, 11, 179–209. [Google Scholar] [CrossRef]
  164. Hoffman, M.R.; Martin, S.T.; Choi, W. Environmental applications of semiconductor photocatalysis. Chem. Rev. 1995, 95, 69–96. [Google Scholar] [CrossRef]
  165. Linsebigler, A.L.; Lu, G.Q.; Yates, J.T. Photocatalysis on TiO2 surfaces: Principles, mechanisms, and selected results. Chem. Rev. 1995, 95, 735–758. [Google Scholar] [CrossRef]
  166. Serpone, N.; Sauve, G.; Koch, R.; Tahiri, H.; Pichat, P.; Piccinini, P.; Pelizzetti, E.; Hidaka, H. Standardization protocol of process efficiencies and activation parameters in heterogeneous photocatalysis: Relative photonic efficiencies ζr. J. Photochem. Photobiol. 1996, 94, 191–203. [Google Scholar] [CrossRef]
  167. Ismail, A.A.; Bahnemann, D.W. Photochemical splitting of water for hydrogen production by photocatalysis: A review. Sol. Energy Mater. Sol. Cells 2014, 28, 85–101. [Google Scholar] [CrossRef]
  168. Wu, H.; Tan, H.L.; Toe, C.Y.; Scott, J.; Wang, L.; Amal, R.; Ng, H.Y. Photocatalytic and photoelectrochemical systems: Similarities and differences. Adv. Mater. 2019, 1–21. [Google Scholar] [CrossRef]
  169. Benck, J.D.; Hellstern, T.R.; Kibsgaard, J.; Chakthranont, P.; Jaramillo, T.F. Catalyzing the hydrogen evolution reaction (HER) with molybdenum sulfide nanomaterials. ACS Catal. 2014, 4, 3957–3971. [Google Scholar] [CrossRef]
  170. Kim, J.S.; Kim, B.; Kim, H.; Kang, K. Amorphous multinary phyllosilicate catalysts for electrochemical water oxidation. Adv. Energy Mater. 2018, 7, 18380–18387. [Google Scholar] [CrossRef]
  171. Hu, C.; Zhang, L.; Gong, J. Recent progress made in the mechanism comprehension and design of electrocatalysts for alkaline water splitting. Energy Environ. Sci. 2019, 12, 2620–2645. [Google Scholar] [CrossRef]
  172. Zheng, J.Y.; Jaroniek, M.; Qiao, S.Z. Design of electrocatalysts for oxygen- and hydrogen-involving energy conversion reactions. Chem. Soc. Rev. 2015, 44, 2060–2086. [Google Scholar] [CrossRef]
  173. Seh, Z.W.; Kibsgaard, J.; Dickens, C.F.; Chorkendorff, I.; Nørskov, J.K.; Jaramillo, T.F. Combining theory and experiment in electrocatalysis: Insights into materials design. Science 2017, 355, 1–12. [Google Scholar] [CrossRef] [Green Version]
  174. Yerga, N.R.M.; Alvarez-Galván, M.C.; Vaquero, F.; Arenales, J.; Fierro, J.L.G. Chapter 3—Hydrogen production from water splitting using photo-semiconductor catalysts. Ren. Hydrog. Technol. Prod. Purif. Storage Appl. Saf. 2013, 43–61. [Google Scholar] [CrossRef]
  175. Li, J.; Wu, N. Semiconductor-based photocatalysts and photoelectrochemical cells for solar fuel generation: A review. Catal. Sci. Technol. 2015, 5, 1360–1384. [Google Scholar] [CrossRef]
  176. Chen, Z.; Dinh, H.; Miller, E. Efficiency Definitions in the Field of PEC. In Photoelectrochemical Water Splitting: Standards, Experimental Methods, and Protocols; Springer: New York, NY, USA, 2013; pp. 7–16. [Google Scholar]
  177. Murphy, A.; Barnes, P.; Randeniya, L.; Plumb, I.; Horne, M.; Glasscock, J. Efficiency of solar water splitting using semiconductor electrodes. Int. J. Hydrogen Energy 2006, 31, 1999–2017. [Google Scholar] [CrossRef]
  178. Smith, W.A. Photoelectrochemical Cell Design, Efficiency, Definitions, Standards, and Protocols. In Photoelectrochemical Solar Fuel Production; Giménez, S., Bisquert, J., Eds.; Springer: Cham, Switzerland, 2016; pp. 163–199. [Google Scholar] [CrossRef]
  179. Fujishima, A.; Honda, K. Electrochemical photolysis of water at a semiconductor electrode. Nature 1972, 238, 37–38. [Google Scholar] [CrossRef]
  180. Li, F.; Liu, Y.; Liu, R.; Sun, Z.; Zhao, D.; Kou, C. Preparation of Ca-doped LaFeO3 nanopowders in a reverse microemulsion and their visible light photocatalytic activity. Mater. Lett. 2010, 64, 223–225. [Google Scholar] [CrossRef]
  181. Wang, G.; Wang, H.; Ling, Y.; Tang, Y.; Yang, X.; Fitzmorris, R.C.; Wang, C.; Zhang, J.Z.; Li, Y. Hydrogen-treated TiO2 nanowire arrays for photoelectrochemical water splitting. Nano Lett. 2011, 11, 3026–3033. [Google Scholar] [CrossRef] [PubMed]
  182. Smith, I.C.; Hoke, E.T.; Solis-Ibarra, D.; McGehee, M.D.; Karunadasa, H.I. A layered hybrid perovskite solar-cell absorber with enhanced moisture stability. Angew. Chem. Int. Ed. 2014, 126, 11414–11417. [Google Scholar] [CrossRef]
  183. Jiang, Q.; Rebollar, D.; Gong, J.; Piacentino, E.L.; Zheng, C.; Xu, T. Pseudohalide-induced moisture tolerance in perovskite CH3NH3Pb(SCN)2I thin films. Angew. Chem. Int. Ed. 2015, 54, 11006. [Google Scholar] [CrossRef] [PubMed]
  184. Da, P.; Cha, M.; Sun, L.; Wu, Y.; Wang, Z.S.; Zheng, G. High-performance perovskite photoanode enabled by Ni passivation and catalysis. Nano Lett. 2015, 15, 3452–3457. [Google Scholar] [CrossRef] [PubMed]
  185. Hoang, M.T.; Pham, N.D.; Han, J.H.; Gardner, J.; Oh, I. Integrated photoelectrolysis of water implemented on organic metal halide perovskite photoelectrod. ACS Appl. Mater. Interfaces 2016, 19, 11904–11909. [Google Scholar] [CrossRef]
  186. Wang, C.; Yang, C.; Wen, T.; Yang, H.G. Surface-functionalized perovskite films for stable photoelectrochemical water splitting. J. Mater. Chem. A 2017, 5, 910–913. [Google Scholar] [CrossRef]
  187. Elseman, A.M.; Selim, M.S.; Luo, L.; Xu, C.Y.; Wang, G.; Jiang, Y.; Liu, D.B.; Liao, L.P.; Hao, Z.; Song, Q. Efficient and stable planar n-i-p perovskite solar cells with negligible hysteresis through solution-processed Cu2O nanocubes as a low-cost hole-transport material. ChemSusChem 2019, 18, 3808–3816. [Google Scholar] [CrossRef]
  188. Upadhyay, S.; Shrivastava, J.; Solanki, A.; Choudhary, S.; Sharma, V.; Kumar, P.; Singh, N.; Satsangi, V.R.; Shrivastav, R.; Waghmare, U.V.; et al. Enhanced photoelectrochemical response of BaTiO3 with Fe doping: Experiments and first-principles analysis. J. Phys. Chem. C 2011, 115/49, 24373–24380. [Google Scholar] [CrossRef]
  189. Jang, J.S.; Ahn, C.W.; Won, S.S.; Kim, J.H.; Choi, W.; Lee, B.S.; Yoon, J.H.; Kim, H.G.; Lee, J.S. Vertically aligned core–shell PbTiO3-TiO2 heterojunction nanotube array for photoelectrochemical and photocatalytic applications. J. Phys. Chem. C 2017, 121, 15063–15070. [Google Scholar] [CrossRef]
  190. Wang, W.; Zhang, W.; Hao, C.; Wu, F.; Liang, Y.; Shi, H.; Wang, J.; Zhang, T.; Hua, Y. Enhanced photoelectrochemical activity and photocatalytic water oxidation of NiO nanoparticle-decorated SrTiO3 nanocube heterostructures: Interaction, interfacial charge transfer and enhanced mechanism. Sol. Energy Mater. Sol. Cells 2016, 152, 1–9. [Google Scholar] [CrossRef]
  191. Wang, F.; Liu, Y.; Ma, Z.; Li, H.; Kang, Z.; Shen, M. Enhanced photoelectrochemical response in SrTiO3 films decorated with carbon quantum dots. New J. Chem. 2013, 37, 290–294. [Google Scholar] [CrossRef]
  192. Takata, T.; Domen, Z. Defect engineering of photocatalysts by doping of aliovalent metal cations for efficient water splitting. J. Phys. Chem. C 2009, 113/45, 19386–19388. [Google Scholar] [CrossRef]
  193. Ikeda, S.; Hirao, K.; Ishino, S.; Matsumura, M.; Ohtani, B. Preparation of platinized strontium titanate covered with hollow silica and its activity for overall water splitting in a novel phase-boundary photocatalytic system. Catal. Today 2006, 117, 343–349. [Google Scholar] [CrossRef] [Green Version]
  194. Kato, H.; Asakura, K.; Kudo, A. Highly efficient water splitting into H2 and O2 over lanthanum-doped NaTaO3 photocatalysts with high crystallinity and surface nanostructure. J. Am. Chem. Soc. 2003, 125, 3082–3089. [Google Scholar] [CrossRef] [PubMed]
  195. Zhang, H.J.; Chen, G.; Li, Y.X.; Teng, Y.J. Electronic structure and photocatalytic properties of copper-doped CaTiO3. Int. J. Hydrogen Energy 2010, 35, 2713–2716. [Google Scholar] [CrossRef]
  196. Li, J.J.; Jia, L.S.; Fang, W.P.; Zeng, J.L. Investigations on the effect of Cu2+/Cu1+ redox couples and oxygen vacancies on photocatalytic activity of treated LaNi1−xCuxO3 (x=0.1, 0.4, 0.5). Int. J. Hydrogen Energy 2010, 35, 5270–5275. [Google Scholar] [CrossRef]
  197. Tang, P.; Chen, H.; Cao, F.; Pan, G. Magnetically recoverable and visible-light-driven nanocrystalline YFeO3 photocatalysts. Catal. Sci. Technol. 2011, 1, 1145–1148. [Google Scholar] [CrossRef]
  198. Díez-García, M.I.; Celorrio, V.; Calvillo, L.; Tiwari, D.; Gómez, R.; Fermín, D.J. YFeO3 photocathodes for hydrogen evolution. Electrochim. Acta 2017, 246, 365–371. [Google Scholar] [CrossRef]
  199. Alexe, M.; Hesse, D. Tip-enhanced photovoltaic effects in bismuth ferrite. Nat. Commun. 2011, 2, 256. [Google Scholar] [CrossRef] [Green Version]
  200. Song, J.; Kim, T.L.; Lee, J.; Cho, S.Y.; Cha, J.; Jeon, S.Y.; An, H.; Kim, W.S.; Jung, Y.S.; Park, J.; et al. Domain-engineered BiFeO3 thin-film photoanodes for highly enhanced ferroelectric solar water splitting. Nano Res. 2018, 11, 642–655. [Google Scholar] [CrossRef]
  201. Zhang, C.; Li, Y.; Chu, M.; Rong, N.; Xiao, P.; Zhang, Y. Hydrogen-treated BiFeO3 nanoparticles with enhanced photoelectrochemical performance. RSC Adv. 2016, 6, 24760–24767. [Google Scholar] [CrossRef]
  202. Moniz, S.J.A.; Blackman, C.S.; Southern, P.; Weaver, P.M.; Tang, J.; Carmalt, C.J. Visible-light driven water splitting over BiFeO3 photoanodes grown via the LPCVD reaction of [Bi(OtBu)3] and [Fe(OtBu)3]2 and enhanced with a surface nickel oxygen evolution catalyst. Nanoscale 2015, 7, 16343–16353. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  203. Gao, F.; Yuan, Y.; Wang, K.; Chen, X.; Chen, F.; Liu, J.; Ren, Z. Preparation and photoabsorption characterization of BiFeO3 nanowires. Appl. Phys. Lett. 2006, 89, 102506–102900. [Google Scholar] [CrossRef]
  204. Joshi, U.A.; Jang, J.S.; Borse, P.H.; Lee, J.S. Microwave synthesis of singlecrystalline perovskite BiFeO3 nanocubes for photoelectrode and photocatalytic applications. Appl. Phys. Lett. 2008, 92, 242106. [Google Scholar] [CrossRef]
  205. Chen, X.Y.; Yu, T.; Gao, F.; Zhang, H.T.; Liu, L.F.; Wang, Y.M.; Li, Z.S.; Zou, Z.G. Application of weak ferromagnetic BiFeO3 films as the photoelectrode material under visible-light irradiation. Appl. Phys. Lett. 2007, 91, 022114. [Google Scholar] [CrossRef]
  206. Cao, D.W.; Wang, Z.J.; Wen, L.Y.; Mi, Y.; Lei, Y. Switchable charge-transfer in the photoelectrochemical energy-conversion process of ferroelectric BiFeO3 photoelectrodes. Angew. Chem. Int. Ed. 2014, 53, 1–6. [Google Scholar] [CrossRef]
  207. Cho, S.; Wang, J.W.; Zhang, W.R.; Suwardi, A.; Wang, H.W.; Wang, D.W.; MacManus-Driscoll, J.L. Single-crystalline thin films for studying intrinsic properties of BiFeO3–SrTiO3 solid solution photoelectrodes in solar energy conversion. Chem. Mater. 2015, 27, 6635–6641. [Google Scholar] [CrossRef]
  208. Rong, N.; Chu, M.; Tang, Y.; Zhang, C.; Cui, Z.; He, H.; Zhang, Y.; Xiao, P. Improved photoelectrocatalytic properties of Ti-doped BiFeO3 films for water oxidation. J. Mater. Sci. 2016, 51, 5712–5723. [Google Scholar] [CrossRef]
  209. Haydous, F.; Scarisoreanu, N.D.; Birjega, R.; Ion, V.; Lippert, T.; Dumitrescu, N.; Moldovan, A.; Andrei, A.; Teodorescu, S.V.; Ghica, C.; et al. Rolling dopant and strain in Y-doped BiFeO 3 epitaxial thin films for photoelectrochemical water splitting. Sci. Rep. 2018, 8, 15826. [Google Scholar] [CrossRef]
  210. Zhang, T.; Shen, Y.; Qiu, Y.; Liu, Y.; Xiong, R.; Shi, J.; Wei, J. Facial synthesis and photoreaction mechanism of BiFeO3/Bi2Fe4O9 heterojunction nanofibers. ACS Sustain. Chem. Eng. 2017, 5, 4630–4636. [Google Scholar] [CrossRef]
  211. Quynh, L.T.; Van, C.N.; Bitla, Y.; Chen, J.W.; Tzeng, W.Y.; Liao, S.C.; Tsai, K.A.; Chen, Y.C.; Wu, C.L.; Lai, C.H.; et al. Self-Assembled BiFeO3-ε-Fe2O3 vertical heteroepitaxy for visible light photoelectrochemistry. Adv. Energy Mater. 2016, 6, 1600686. [Google Scholar] [CrossRef]
  212. Khoomortezaei, S.; Abdizadeh, H.; Golobostanfard, M.R. Triple layer heterojunction WO3/BiVO4/BiFeO3 porous photoanode for efficient photoelectrochemical water splitting. ACS Appl. Energy Mater. 2019, 2, 6428–6439. [Google Scholar] [CrossRef]
  213. Li, S.; Jing, L.; Fu, W.; Yang, L.; Xin, B.; Fu, H. Photoinduced charge property of nanosized perovskite-type LaFeO3 and its relationships with photocatalytic activity under visible irradiation. Mater. Res. Bull. 2007, 42, 203–212. [Google Scholar] [CrossRef]
  214. Celorrio, V.; Bradley, K.; Weber, O.J.; Hall, S.R.; Fermin, D.J. Photoelectrochemical properties of LaFeO3 nanoparticles. ChemElectroChem 2014, 1, 1667–1671. [Google Scholar] [CrossRef]
  215. Martinelli, G.; Carotta, M.C.; Ferroni, M.; Sadaoka, Y.; Traversa, Y. Screen-printed perovskite-type thick films as gas sensors for environmental monitoring. Sens. Act. B Chem. 1999, 55, 99–110. [Google Scholar] [CrossRef]
  216. Parida, K.M.; Reddy, K.H.; Martha, S.; Das, D.P.; Biswal, N. Fabrication of nanocrystalline LaFeO3: An efficient sol–gel auto-combustion assisted visible light responsive photocatalyst for water decomposition. Int. J. Hydrogen Energy 2010, 35, 12161–12168. [Google Scholar] [CrossRef]
  217. Sora, I.N.; Fontana, F.; Passalacqua, R.; Ampelli, C.; Perathoner, S.; Centi, G.; Parrino, F.; Palmisano, L. Photoelectrochemical properties of doped lanthanum orthoferrites. Electrochim. Acta 2013, 109, 710–715. [Google Scholar] [CrossRef] [Green Version]
  218. Yu, Q.; Meng, X.; Wang, T.; Li, P.; Liu, L.; Chang, K.; Liu, G.; Ye, J. A highly durable p-LaFeO3/n-Fe2O3 photocell for effective water splitting under visible light. Chem. Commun. 2015, 51, 3630–3633. [Google Scholar] [CrossRef]
  219. May, K.J.; Fenning, D.P.; Ming, T.; Hong, W.T.; Lee, D.; Stoerzinger, K.A.; Biegalski, M.D.; Kolpak, A.M.; Horn, Y.S. Thickness-dependent photoelectrochemical water splitting on ultrathin LaFeO3 films grown on Nb:SrTiO3. J. Phys. Chem. Lett. 2015, 6, 977–985. [Google Scholar] [CrossRef]
  220. Peng, Q.; Shan, B.; Wen, Y.; Chen, R. Enhanced charge transport of LaFeO3 via transition metal (Mn, Co, Cu) doping for visible light photoelectrochemical water oxidation. Int. J. Hydrogen Energy 2015, 40, 15423–15431. [Google Scholar] [CrossRef]
  221. Ji, W.; Yao, K.; Lim, Y.F.; Liang, Y.C.; Suwardi, A. Epitaxial ferroelectric BiFeO3 thin films for unassisted photocatalytic water splitting. Appl. Phys. Lett. 2013, 103, 062901. [Google Scholar] [CrossRef]
  222. Wu, X.; Li, H.; Wang, X.; Jiang, L.; Xi, J.; Du, G.; Ji, Z. Ferroelectric enhanced photoelectrochemical water splitting in BiFeO3/TiO2 composite photoanode. J. Alloys Compd. 2018, 783, 643–651. [Google Scholar] [CrossRef]
  223. Wheeler, G.P.; Choi, K.S. Photoelectrochemical properties and stability of nanoporous p-type LaFeO3 photoelectrodes prepared by electrodeposition. ACS Energy Lett. 2017, 2, 2378–2382. [Google Scholar] [CrossRef]
  224. Pawar, G.S.; Tahir, A.A. Unbiased spontaneous solar fuel production using stable LaFeO3 photoelectrode. Sci. Rep. 2018, 8, 3501. [Google Scholar] [CrossRef] [PubMed]
  225. Nassar, I.M.; Wu, S.; Li, L.; Li, X. Facile preparation of n-type LaFeO3 perovskite film for efficient photoelectrochemical water splitting. ChemistrySelect 2018, 3, 968–972. [Google Scholar] [CrossRef]
  226. Pawar, G.S.; Elikkottil, A.; Seetha, S.; Reddy, K.S.; Pesala, B.; Tahir, A.A.; Mallick, T.K. Enhanced photoactivity and hydrogen generation of LaFeO3 photocathode by plasmonic silver nanoparticle incorporation. ACS Appl. Energy Mater. 2018, 1, 3449–3456. [Google Scholar] [CrossRef]
  227. Wang, P.; He, Y.; Mi, Y.; Zhu, J.; Zhang, F.; Liu, Y.; Yang, Y.; Chen, M.; Cao, D. Enhanced photoelectrochemical performance of LaFeO3 photocathode with Au buffer layer. RSC Adv. 2019, 9, 26780–26786. [Google Scholar] [CrossRef] [Green Version]
  228. Yang, J.; Wang, D.; Han, X.; Li, C. Roles of cocatalysts in photocatalysis and photoelectrocatalysis. Acc. Chem. Res. 2013, 46, 1900–1909. [Google Scholar] [CrossRef]
  229. Liu, Q.; Zhou, Y.; You, L.; Wang, J.; Shen, M.; Fang, L. Enhanced ferroelectric photoelectrochemical properties of polycrystalline BiFeO3 film by decorating with Ag nanoparticles. Appl. Phys. Lett. 2016, 108, 022902. [Google Scholar] [CrossRef]
  230. Ren, Y.; Nan, F.; You, L.; Zhou, Y.; Wang, Y.; Wang, J.; Su, X.; Shen, M.; Fang, L. Enhanced photoelectrochemical performance in reduced graphene oxide/BiFeO3 heterostructures. Small 2017, 13, 1603457. [Google Scholar] [CrossRef]
  231. Herzog, H.; Golomb, D. Carbon capture and storage from fossil fuel use. Encycl. Energy. 2004, 1, 277–287. [Google Scholar] [CrossRef]
  232. Lee, J.H.; Trimm, D.L. Catalytic combustion of methane. Fuel Process. Technol. 1995, 42, 339–359. [Google Scholar] [CrossRef]
  233. Campagnoli, E.; Tavares, A.C.; Fabbrini, L.; Rossetti, I.; Dubitsky, Y.A.; Zaopo, A.; Forni, L. La1−xA′xCo1−yFeyO3±δ (A′ = Ce,Sr) catalysts for the flameless combustion of methane. J. Mater. Sci. 2006, 41, 4713–4719. [Google Scholar] [CrossRef]
  234. Chen, J.; Arandiyan, H.; Gao, X.; Li, J. Recent Advances in Catalysts for Methane Combustion. Catal. Surv. Asia 2015, 19, 140–171. [Google Scholar] [CrossRef]
  235. Jin, J.; Li, C.; Tsang, C.W.; Xu, B.; Lian, C. Catalytic combustion of methane over Pt–Ce oxides under scarce oxygen condition. Ind. Eng. Chem. Res. 2016, 55, 2293–2301. [Google Scholar] [CrossRef]
  236. Eriksson, S.; Wolf, M.; Scheider, A.; Mantzaras, J.; Raimondi, F.; Boutonnet, M.; Jaras, S. Fuel-rich catalytic combustion of methane in zero emissions power generation processes. Catal. Today. 2006, 117, 447–453. [Google Scholar] [CrossRef]
  237. Ciuparu, D.; Lyubovsky, M.R.; Altman, E.; Pfefferle, L.D.; Datye, A. Catalytic combustion of methane over palladium-based catalysts. Catal. Rev. 2002, 44, 593–649. [Google Scholar] [CrossRef]
  238. Cimino, S.; Colonna, S.; De Rossi, S.; Faticanti, M.; Lisi, L.; Pettiti, I.; Porta, P. Methane Combustion and CO oxidation on zirconia-supported La, Mn oxides and LaMnO3 perovskite. J. Catal. 2002, 205, 309–317. [Google Scholar] [CrossRef]
  239. Landi, G.; Barbato, P.S.; Di Sarli, V.; Di Benedetto, A. Multifuel Catalytic Combustion in the Presence of Carbon Dioxide over Fully and Partially Perovskite-Coated Monoliths. Ind. Eng. Chem. Res. 2017, 56, 4920–4928. [Google Scholar] [CrossRef] [Green Version]
  240. Barbato, P.S.; Di Sarli, V.; Landi, G.; Di Benedetto, A. High pressure methane catalytic combustion over novel partially coated LaMnO3-based monoliths. Chem. Eng. J. 2015, 259, 381–390. [Google Scholar] [CrossRef]
  241. Di Sarli, V.; Barbato, P.S.; Di Benedetto, A.; Landi, G. Start-up behavior of a LaMnO3 partially coated monolithic combustor at high pressure. Catal. Today 2015, 242, 200–210. [Google Scholar] [CrossRef]
  242. Landi, G.; Di Benedetto, A.; Barbato, P.S.; Russo, G.; Di Sarli, V. Transient behavior of structured LaMnO3 catalyst during methane combustion at high pressure. Chem. Eng. Sci. 2014, 116, 350–358. [Google Scholar] [CrossRef]
  243. Barbato, P.S.; Di Benedetto, A.; Di Sarli, V.; Landi, G.; Pirone, R. High-Pressure Methane Combustion over a Perovskyte Catalyst. Ind. Eng. Chem. Res. 2012, 51, 7547–7558. [Google Scholar] [CrossRef]
  244. Nakamura, T.; Petzow, G.; Gauckler, L.J. Stability of the perovskite phase LaBO3 (B = V, Cr, Mn, Fe, Co, Ni) in reducing atmosphere I. Experimental results. Mater. Res. Bull. 1979, 14, 649–659. [Google Scholar] [CrossRef]
  245. Spinicci, R.; Tofanari, A.; Delmasto, A.; Mazza, D.; Ronchetti, S. Catalytic properties of stoichiometric and non-stoichiometric LaFeO3 perovskite for total oxidation of methane. Mater. Chem. Phys. 2002, 76, 20–25. [Google Scholar] [CrossRef]
  246. Alifanti, M.; Florea, M.; Cortes-Corberan, V.; Endruschat, U.; Delmon, B.; Parvulescu, V.I. Effect of LaCoO3 perovskite deposition on ceria-based supports on total oxidation of VOC. Catal. Today. 2006, 112, 169–173. [Google Scholar] [CrossRef]
  247. Popescu, I.; Urda, A.; Yuzhakova, T.; Marcu, I.C.; Kovacs, J.; Sandulescu, I. BaTiO3 and PbTiO3 perovskite as catalysts for methane combustion. Comptes Rendus Chim. 2009, 12, 1072–1078. [Google Scholar] [CrossRef]
  248. Taguchi, H.; Nakade, K.; Yosinaga, M.; Kato, M.; Hirota, K. Methane oxidation on perovskite-type Ca(Mn1−xTix)O3−δ. J. Am. Ceram. Soc. 2008, 310, 2007–2009. [Google Scholar] [CrossRef]
  249. Tofan, C.; Klvana, D.; Kichnerova, J. Decomposition of nitric oxide over perovskite oxide catalysts: Effect of CO2, H2O and CH4. Appl. Catal. B 2002, 36, 311–323. [Google Scholar] [CrossRef]
  250. Najjar, H.; Lamonier, J.; Mentré, O.; Giraudon, J.; Baris, H. Combustion synthesis of LaMn1−xAlxO3+δ (0 ≤ x ≤ 1): Tuning catalytic properties for methane deep oxidation. Catal. Sci. Technol. 2013, 3, 1002–1016. [Google Scholar] [CrossRef]
  251. Miao, F.; Wang, F.; Mao, D.; Guo, X.; Yu, J.; Huang, H.; Lu, G. Effect of different reaction conditions on catalytic activity of La((Mn, Fe)O3+λ catalyst for methane combustion. Mater. Res. Express. 2019, 6, 5001. [Google Scholar] [CrossRef]
  252. Edokpayi, J.N.; Odiyo, J.O.; Durowoju, O.S. Impact of Wastewater on Surface Water Quality in Developing Countries: A Case Study of South Africa. In Water Quality; Hlanganani, T., Ed.; IntechOpen: London, UK, 2017. [Google Scholar] [CrossRef] [Green Version]
  253. Oh, H.J.; Lee, J.H.; Kim, Y.J.; Suh, S.J.; Lee, J.H.; Chi, C.S. Synthesis of effective titania nanotubes for wastewater purification. Appl. Catal. B 2008, 84, 142–147. [Google Scholar] [CrossRef]
  254. Manu, B.; Chaudhari, S. Anaerobic decolorisation of simulated textile wastewater containing azo dyes. Bioresour. Technol. 2002, 82, 225–231. [Google Scholar] [CrossRef]
  255. Lai, C.W.; Juan, J.C.; Ko, W.B.; Bee Abd Hamid, S. An overview: Recent development of titanium oxide nanotubes as photocatalyst for dye degradation. Int. J. Photoenergy 2014, 2014, 1–14. [Google Scholar] [CrossRef]
  256. Ali, I.; Gupta, V.K. Advances in water treatment by adsorption technology. Nat. Protoc. 2006, 1, 2661–2667. [Google Scholar] [CrossRef] [PubMed]
  257. Aoudj, S.; Drouiche, N.; Hecini, M.; Ouslimane, T.; Palaouane, B. Coagulation as a post-treatment method for the defluoridation of photovoltaic cell manufacturing wastewater. Procedia Eng. 2012, 33, 111–120. [Google Scholar] [CrossRef] [Green Version]
  258. Kim, T.H.; Park, C.; Shin, E.B.; Kim, S. Decolorization of disperse and reactive dyes by continuous electrocoagulation process. Desalination 2002, 150, 165–175. [Google Scholar] [CrossRef]
  259. Apin, R.; Waite, T.D. Comparison of three advanced oxidation processes for degradation of textile dyes. Water Sci. Technol. 2000, 42, 345–354. [Google Scholar] [CrossRef]
  260. Cham, K.; Seok, J.D.; Se, G.L.; Sung, J.L.; Ho, Y.K. Visible-light absorptivity of a zincoxysulfide (ZnOxS1-x) composite semiconductor and its photocatalytic activities for degradation of organic pollutants under visible-light irradiation. Appl. Catal. A 2007, 330, 127–133. [Google Scholar] [CrossRef]
  261. Yu, J.; Wang, Y.; Xiao, W. Enhanced photoelectrocatalytic performance of SnO2/TiO2 rutile composite films. J. Mater. Chem. A 2013, 1, 10727–10735. [Google Scholar] [CrossRef]
  262. Vassilis, J.I.; Stavros, G.P. Chapter 4. Adsorption and Ion Exchange. In Adsorption, Ion Exchange and Catalysis, 1st ed.; Poulopoulos, S., Inglezakis, V., Eds.; Elsevier: London, UK, 2006; pp. 243–353. [Google Scholar]
  263. Geçgel, Ü.; Özcan, G.; Gürpınar, G.Ç. Removal of methylene blue from aqueous solution by activated carbon prepared from Pea Shells (Pisum sativum). J. Chem. 2013, 1–9. [Google Scholar] [CrossRef] [Green Version]
  264. Kuang, Y.; Zhang, X.; Zhou, S. Adsorption of methylene blue in water onto activated carbon by surfactant modification. Water 2020, 12, 587. [Google Scholar] [CrossRef] [Green Version]
  265. Bhargavi, R.J.; Maheshwari, U.; Gupta, S. Synthesis and use of alumina nanoparticles as an adsorbent for the removal of Zn(II) and CBG dye from wastewater. Int. J. Ind. Chem. 2014, 6, 31–41. [Google Scholar] [CrossRef] [Green Version]
  266. Ali, S.; Abbas, Y.; Zuhra, Z.; Butler, I.S. Synthesis of γ-alumina (Al2O3) nanoparticles and their potential for use as an adsorbent in the removal of methylene blue dye from industrial wastewater. Nanoscale Advances. 2018, 1, 213–218. [Google Scholar] [CrossRef] [Green Version]
  267. Tahir, S.S.; Rauf, N. Removal of a cationic dye from aqueous solutions by adsorption onto bentonite clay. Chemosphere 2006, 63, 1842–1848. [Google Scholar] [CrossRef]
  268. Hu, Q.H.; Qiao, S.Z.; Haghseresht, F.; Wilson, M.A.; Lu, G.Q. Adsorption study for removal of basic red dye using bentonite. Ind. Eng. Chem. Res. 2006, 45, 733–738. [Google Scholar] [CrossRef]
  269. Wang, S.; Li, H.; Xu, L. Application of zeolite MCM-22 for basic dye removal from wastewater. J. Colloid Interface Sci. 2006, 295, 71–78. [Google Scholar] [CrossRef]
  270. Barrera-Díaz, C.E.; Balderas-Hernández, P.; Bilyeu, B. Electrocoagulation: Fundamentals and prospectives. Electrochem. Water Wastew. Treatm. 2018, 4, 61–76. [Google Scholar] [CrossRef]
  271. Bhandari, V.M.; Ranade, V.V. Advanced Physico-Chemical Methods of Treatment for Industrial. Ind. Wastew. Treatm. Recycl. Reuse Wastew. 2014, 81–140. [Google Scholar] [CrossRef]
  272. Ghosh, D.; Solanki, H.; Purkait, M.K. Removal of Fe(II) from tap water by electrocoagulation technique. J. Hazard. Mater. 2008, 155, 135–143. [Google Scholar] [CrossRef]
  273. Moussa, D.T.; El-Naas, M.H.; Nasser, M.; Al-Marri, M.J. A comprehensive review of electrocoagulation for water treatment: Potentials and challenges. J. Environ. Manag. 2017, 186, 24–41. [Google Scholar] [CrossRef] [PubMed]
  274. Mustafa, M.G. Biochemical basis of ozone toxicity. Free Radical Biol. Med. 1990, 9, 245–265. [Google Scholar] [CrossRef]
  275. Wilhelm, P.; Stephan, D. Photodegradation of rhodamine B in aqueous solution via SiO2@TiO2 nano-spheres. J. Photochem. Photobiol. A 2007, 185, 19–25. [Google Scholar] [CrossRef]
  276. Watanabe, T.; Takizawa, T.; Honda, K. Photocatalysis through excitation of adsorbates. 1. Highly efficient N-deethylation of rhodamine B adsorbed to cadmium sulfide. J. Phys. Chem. 1977, 81, 1845–1851. [Google Scholar] [CrossRef]
  277. Takizawa, T.; Watanabe, K.; Honda, J. Photocatalysis through excitation of adsorbates. 2. A comparative study of Rhodamine B and methylene blue on cadmium sulfide. J. Phys. Chem. 1978, 82, 1391–1396. [Google Scholar] [CrossRef]
  278. Qu, P.; Zhao, J.; Shen, T.; Hidaka, H. TiO2-assisted photodegradation of dyes: A study of two competitive primary processes in the degradation of RB in an aqueous TiO2 colloidal solution. J. Mol. Catal. A 1998, 129, 257–268. [Google Scholar] [CrossRef]
  279. Haque, E.; Jun, J.W.; Jhung, S.H. Adsorptive removal of methyl orange and methylene blue from aqueous solution with a metal-organic framework material, iron terephthalate (MOF-235). J. Hazard. Mater. 2011, 185, 507–511. [Google Scholar] [CrossRef]
  280. Li, Y.; Li, X.; Li, J.; Yin, J. Photocatalytic degradation of methyl orange by TiO2-coated activated carbon and kinetic study. Water Res. 2006, 40, 1119–1126. [Google Scholar] [CrossRef]
  281. Qian, X.; Chen, Z.; Yang, X.; Zhao, W.; Liu, C.; Sun, T.; Zhou, D.; Yang, Q.; Wei, G.; Fan, M. Perovskite cesium lead bromide quantum dots: A new efficient photocatalyst for degrading antibiotic residues in organic system. J. Clean. Prod. 2019, 249, 119335. [Google Scholar] [CrossRef]
  282. Chen, C.; Zhou, J.; Geng, J.; Bao, R.; Wang, Z.; Xia, J.; Li, H. Perovskite LaNiO3/TiO2 step-scheme heterojunction with enhanced photocatalytic activity. Appl. Surf. Sci. 2019, 503, 144287. [Google Scholar] [CrossRef]
  283. Bresolin, B.M.; Hammouda, S.B.; Sillanpää, M. An emerging visible-light organic–inorganic hybrid perovskite for photocatalytic applications. Nanomaterials 2020, 10, 115. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  284. Elmorsi, T.M. Equilibrium isotherms and kinetic studies of removal of methylene blue dye by adsorption onto miswak leaves as a natural adsorbent. J. Environ. Prot. 2011, 2, 817–827. [Google Scholar] [CrossRef] [Green Version]
  285. Shah, N.S.; Khan, J.A.; Sayed, M.; Khan, Z.U.H.; Iqbal, J.; Arshad, S.; Junaid, M.; Khan, H.M. Synergistic effects of H2O2 and S2O82– in the gamma radiation induced degradation of Congo-red dye: Kinetics and toxicities evaluation. Sep. Purif. Technol. 2019, 233, 115966. [Google Scholar] [CrossRef]
  286. Chaibakhsh, N.; Ahmadi, N.; Zanjanchi, M.A. Optimization of photocatalytic degradation of neutral red dye using TiO2 nanocatalyst via Box-Behnken design. Desalinat. Water Treat. 2015, 57, 9296–9306. [Google Scholar] [CrossRef]
  287. Wahab, H.S.; Hussain, A.A. Photocatalytic oxidation of phenol red onto nanocrystalline TiO2 particles. J. Nanostruct. Chem. 2016, 6, 261–274. [Google Scholar] [CrossRef] [Green Version]
  288. Abdollahi, Y.; Abdullah, A.H.; Zainal, Z.; Yusof, N.A. Photocatalytic degradation of p-cresol by zinc oxide under UV Irradiation. Int. J. Mol.Sci. 2011, 13, 302–315. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  289. Li, P.; Liu, C.; Wu, G.; Heng, Y.; Lin, S.; Ren, A.; Lv, K.; Xiao, L.; Shi, W. Solvothermal synthesis and visible light-driven photocatalytic degradation for tetracycline of Fe-doped SrTiO3. RSC Adv. 2014, 4, 47615–47624. [Google Scholar] [CrossRef]
  290. Karstens, T.; Kobs, K. Rhodamine B and rhodamine 101 as reference substances for fluorescence quantum yield measurements. J. Phys. Chem. 1980, 84, 1871–1872. [Google Scholar] [CrossRef]
  291. Szeto, W.; Kan, C.W.; Yuen, C.W.M.; Chan, S.W.; Lam, K.H. Effective photodegradation of methyl orange using fluidized bed reactor loaded with cross-linked chitosan embedded nano-CdS photocatalyst. Int. J. Chem. Eng. 2014, 2014, 1–16. [Google Scholar] [CrossRef]
  292. Paşka, O.; Ianoş, R.; Păcurariu, C.; Brădeanu, A. Magnetic nanopowder as effective adsorbent for the removal of Congo Red from aqueous solution. Water Sci. Technol. 2013, 69, 1234–1240. [Google Scholar] [CrossRef]
  293. Alnuaimi, M.M.; Rauf, M.A.; Ashraf, S.S. Comparative decoloration study of Neutral Red by different oxidative processes. Dyes Pigments. 2007, 72, 367–371. [Google Scholar] [CrossRef]
  294. Fiege, H. Cresols and Xylenols. In Ullmann’s Encyclopedia of Industrial Chemistry; Wiley: Hoboken, NJ, USA, 2000. [Google Scholar] [CrossRef]
  295. Cosentino, U.; Pitea, D.; Moro, G.; Saracino, G.A.A.; Caria, P.; Varì, R.M.; Colombo, L.; Forloni, G.; Tagliavini, F.; Salmona, M. The anti-fibrillogenic activity of tetracyclines on PrP 106–126: A 3D-QSAR study. J. Mol. Model. 2008, 14, 987–994. [Google Scholar] [CrossRef] [PubMed]
  296. Primo, A.; Corma, A.; García, H. Titania supported gold nanoparticles as photocatalyst. Phys. Chem. Chem. Phys. 2011, 13, 886–910. [Google Scholar] [CrossRef] [PubMed]
  297. Zhao, J.C.; Wu, T.X.; Wu, K.Q.; Oikawa, K.; Hidaka, H.; Serpone, N. Photoassisted degradation of dye pollutants (III) degradation of the cationic dye rhodamine B in aqueous anionic surfactant/TiO2 dispersions under visible light irradiation: Evidence for the need of substrate adsorption on TiO2 particles. Environ. Sci. Technol. 1998, 32, 2394–2400. [Google Scholar] [CrossRef]
  298. Kominami, H.; Kumamoto, H.; Kera, Y.; Ohtani, B. Immobilization of highly active titanium(IV) oxide particles—A novel strategy of preparation of transparent photocatalytic coatings. Appl. Catal. B 2001, 30, 329–335. [Google Scholar] [CrossRef] [Green Version]
  299. Shankar, K.; Basham, J.I.; Allam, N.K.; Varghese, O.K.; Mor, G.K.; Feng, X.; Paulose, M.; Seabold, J.A.; Choi, K.S.; Grimes, C.A. Recent advances in the use of TiO2 nanotube and nanowire arrays for oxidative photoelectrochemistry. J. Phys. Chem. C 2009, 113, 6327–6359. [Google Scholar] [CrossRef]
  300. Yan, G.; Zheng, L.; Xie, L.; Weng, X.; Ye, J. Nature of Ag-Bi-codoped TiO2 visible light photocatalyst. Rare Metals. 2011, 30, 259–266. [Google Scholar] [CrossRef]
  301. Wei, X.; Xu, G.; Ren, Z.; Xu, C.; Weng, W.; Shen, G.; Han, G. Single-crystal-like mesoporous SrTiO3 spheres with enhanced photocatalytic performance. J. Am. Ceram. Soc. 2010, 93, 1297–1305. [Google Scholar] [CrossRef]
  302. Da Silva, L.F.; Avansi, W.; Andrés, J.; Ribeiro, C.; Moreira, M.L.; Longo, E.; Mastelaro, V.R. Long-range and short-range structures of cube-like shape SrTiO3 powders: Microwave-assisted hydrothermal synthesis and photocatalytic activity. Phys. Chem. Chem. Phys. 2013, 15, 12386. [Google Scholar] [CrossRef]
  303. Puangpetch, T.; Sreethawong, T.; Yoshikawa, S.; Chavadej, S. Synthesis and photocatalytic activity in methyl orange degradation of mesoporous-assembled SrTiO3 nanocrystals prepared by sol–gel method with the aid of structure-directing surfactant. J. Mol. Catal. A Chem. 2008, 287, 70–79. [Google Scholar] [CrossRef]
  304. Xie, T.H.; Sun, X.; Lin, J. Enhanced photocatalytic degradation of RhB driven by visible light-induced MMCT of Ti(IV)−O−Fe(II) Formed in Fe-Doped SrTiO3. J. Phys. Chem. C 2008, 112, 9753–9759. [Google Scholar] [CrossRef]
  305. Dong, B.; Li, Z.; Li, Z.; Xu, X.; Song, M.; Zheng, W.; Wang, C.; Al-Deyab, S.S.; El-Newehy, M. Highly efficient LaCoO3 nanofibers catalysts for photocatalytic degradation of Rhodamine B. J. Am. Ceram. Soc. 2010, 93, 3587–3590. [Google Scholar] [CrossRef]
  306. Jung, W.Y.; Hong, S.S. Synthesis of LaCoO3 nanoparticles by microwave process and their photocatalytic activity under visible light irradiation. J. Ind. Eng. Chem. 2013, 19, 157–160. [Google Scholar] [CrossRef]
  307. Gao, F.; Chen, X.Y.; Yin, K.B.; Dong, S.; Ren, Z.F.; Yuan, F.; Yu, T.; Zou, Z.; Liu, J.M. Visible-light photocatalytic properties of weak magnetic BiFeO3 nanoparticles. Adv. Mater. 2007, 19, 2889–2892. [Google Scholar] [CrossRef]
  308. Mohan, S.; Subramanian, B.; Bhaumik, I.; Gupta, P.K.; Jaisankar, S.N. Nanostructured Bi(1-x)Gd(x)FeO3—A multiferroic photocatalyst on its sunlight driven photocatalytic activity. RSC Adv. 2014, 4, 16871–16878. [Google Scholar] [CrossRef]
  309. Guo, R.; Fang, L.; Dong, W.; Zheng, F.; Shen, M. Enhanced photocatalytic activity and ferromagnetism in Gd doped BiFeO3 nanoparticles. J. Phys. Chem. C 2010, 114, 21390–21396. [Google Scholar] [CrossRef]
  310. Feng, Y.N.; Wang, H.C.; Luo, Y.D.; Shen, Y.; Lin, Y.H. Ferromagnetic and photocatalytic behaviors observed in Ca-doped BiFeO3 nanofibres. J. Appl. Phys. 2013, 113, 146101. [Google Scholar] [CrossRef]
  311. Zhang, Z.A.; Liu, H.Y.; Lin, Y.H.; Wei, Y.; Nan, C.W.; Deng, X.L. Influence of La doping on magnetic and optical properties of bismuth ferrite nanofibers. J. Nanomater. 2012, 2012, 1–5. [Google Scholar] [CrossRef] [Green Version]
  312. Ramezanalizadeh, H.; Manteghi, F. Design and development of a novel BiFeO3/CuWO4 heterojunction with enhanced photocatalytic performance for the degradation of organic dyes. J. Photoch. Photobio. A 2017, 338, 60–71. [Google Scholar] [CrossRef]
  313. Liu, Y.; Ding, S.; Xu, J.; Zhang, H.; Yang, S.; Duan, X.; Sun, H.; Wang, S. Preparation of a p-n heterojunction BiFeO3@TiO2 photocatalyst with a core–shell structure for visible-light photocatalytic degradation. Chin. J. Catal. 2017, 38, 1052–1062. [Google Scholar] [CrossRef]
  314. Deganello, F.; Tummino, M.L.; Calabrese, C.; Testa, M.L.; Avetta, P.; Fabbri, D.; Prevot, A.B.; Montoneri, E.; Magnacca, G. A new, sustainable LaFeO3 material prepared from biowaste-sourced soluble substances. New J. Chem. 2015, 39, 877–885. [Google Scholar] [CrossRef]
  315. Li, L.; Wang, X.; Zhang, Y. Enhanced visible light-responsive photocatalytic activity of LnFeO3 (Ln=La, Sm) nanoparticles by synergistic catalysis. Mater. Res. Bull. 2014, 50, 18–22. [Google Scholar] [CrossRef]
  316. Thirumalairajan, S.; Girija, K.; Ganesh, I.; Mangalaraj, D.; Viswanathan, C.; Balamurugan, A.; Ponpandian, N. Controlled synthesis of perovskite LaFeO3 microsphere composed of nanoparticles via self-assembly process and their associated photocatalytic activity. Chem. Eng. J. 2012, 209, 420–428. [Google Scholar] [CrossRef]
  317. Shen, H.; Xue, T.; Wang, Y.; Cao, G.; Lu, Y.; Fang, G. Photocatalytic property of perovskite LaFeO3 synthesized by sol-gel process and vacuum microwave calcination. Mater. Res. Bull. 2016, 84, 15–24. [Google Scholar] [CrossRef]
  318. Thirumalairajan, S.; Girija, K.; Hebalkar, N.Y.; Mangalaraj, D.; Viswanathan, C.; Ponpandian, N. Shape evolution of perovskite LaFeO3 nanostructures: A systematic investigation of growth mechanism, properties and morphology dependent photocatalytic activities. RSC Adv. 2013, 3, 7549–7561. [Google Scholar] [CrossRef]
  319. Wei, Z.X.; Wang, Y.; Liu, J.P.; Xiao, C.M.; Zeng, W.W. Synthesis, magnetization and photocatalytic activity of LaFeO3 and LaFe0.5Mn0.5-xO3-δ. Mater. Chem. Phys. 2012, 136, 755–761. [Google Scholar] [CrossRef]
  320. Liang, Q.; Jin, J.; Liu, C.; Xu, S.; Li, Z. Constructing a novel p-n heterojunction photocatalyst LaFeO3/g-C3N4 with enhanced visible-light-driven photocatalytic activity. J. Alloys Compd. 2017, 709, 542–548. [Google Scholar] [CrossRef]
  321. Ye, Y.; Yang, H.; Wang, X.; Feng, W. Photocatalytic, Fenton and photo-Fenton degradation of RhB over Z-scheme g-C3N4/LaFeO3 heterojunction photocatalysts. Mat. Sci. Semicon. Proc. 2018, 82, 14–24. [Google Scholar] [CrossRef]
  322. Wu, Y.; Wang, H.; Tu, W.; Liu, Y.; Tan, Y.Z.; Yuan, X.; Chew, J.W. Quasi-polymeric construction of stable perovskite-type LaFeO3/g-C3N4 heterostructured photocatalyst for improved Z-scheme photocatalytic activity via solid p-n heterojunction interfacial effect. J. Hazard. Mater. 2018, 347, 412–422. [Google Scholar] [CrossRef]
  323. Zou, F.; Jiang, Z.; Qin, X.; Zhao, Y.; Jiang, L.; Zhi, J.; Xiao, T.; Edwards, P.P. Template-free synthesis of mesoporous N-doped SrTiO3 perovskite with high visible-light-driven photocatalytic activity. Chem. Commun. 2012, 48, 8514–8516. [Google Scholar] [CrossRef] [Green Version]
  324. Jiang, L.; Qui, Y.; Yi, Z. Potassium niobate nanostructures: Controllable morphology, growth mechanism, and photocatalytic activity. J. Mater. Chem. A 2013, 1, 2878–2885. [Google Scholar] [CrossRef]
  325. Li, G.; Yi, Z.; Bai, Y.; Zhang, W.; Zhang, H. Anisotropy in photocatalytic oxidization activity of NaNbO3 photocatalyst. Dalton Trans. 2012, 41, 10194–10198. [Google Scholar] [CrossRef] [PubMed]
  326. Ding, J.; Lü, X.; Shu, H.; Xie, J.; Zhang, H. Microwave-assisted synthesis of perovskite ReFeO3 (Re: La, Sm, Eu, Gd) photocatalyst. Mat. Sci. Eng. B 2010, 171, 31–34. [Google Scholar] [CrossRef]
  327. Cho, C.M.; Noh, J.H.; Cho, I.S.; An, J.S.; Hong, K.S.; Kim, J.Y. Low-temperature hydrothermal synthesis of pure BiFeO3 nanopowders using triethanolamine and their applications as visible-light photocatalysts. J. Am. Ceram. Soc. 2008, 91, 3753–3755. [Google Scholar] [CrossRef]
  328. Huo, Y.; Miao, M.; Zhang, Y.; Zhu, J.; Li, H. Aerosol-spraying preparation of a mesoporous hollow spherical BiFeO3 visible photocatalyst with enhanced activity and durability. Chem. Commun. 2011, 47, 2089–2091. [Google Scholar] [CrossRef] [PubMed]
  329. Hao, C.; Wen, F.; Xiang, J.; Hou, H.; Lv, W.; Lv, Y.; Hu, W.; Liu, Z. Photocatalytic performances of BiFeO3 particles with the average size in nanometer, submicrometer, and micrometer. Mater. Res. Bull. 2014, 50, 369–373. [Google Scholar] [CrossRef]
  330. Li, S.; Zhao, Y.; Wang, C.; Li, D.; Gao, K. Fabrication and characterization unique ribbon-like porous Ag/LaFeO3 nanobelts photocatalyst via electrospinning. Mater. Lett. 2016, 170, 122–125. [Google Scholar] [CrossRef]
  331. Fu, S.; Niu, H.; Tao, Z.; Song, J.; Mao, C.; Zhang, S.; Chen, C.; Wang, D. Low temperature synthesis and photocatalytic property of perovskite-type LaCoO3 hollow spheres. J. Alloys Compd. 2013, 576, 5–12. [Google Scholar] [CrossRef]
  332. Yan, L.; Zhang, T.; Lei, W.; Xu, Q.; Zhou, X.; Xu, P.; Wang, Y.; Liu, G. Catalytic activity of gold nanoparticles supported on KNbO3 microcubes. Catal. Today. 2014, 224, 140–146. [Google Scholar] [CrossRef]
  333. Dhanalakshmi, R.; Muneeswaran, M.; Shalini, K.; Giridharan, N.V. Enhanced photocatalytic activity of La-substituted BiFeO3 nanostructures on the degradation of phenol red. Mater. Lett. 2016, 165, 205–209. [Google Scholar] [CrossRef]
  334. Hou, L.; Sun, G.; Liu, K.; Li, Y.; Gao, F. Preparation, characterization and investigation of catalytic activity of Li-doped LaFeO3 nanoparticles. J. Sol Gel Sci. Techn. 2006, 40, 9–14. [Google Scholar] [CrossRef]
  335. Li, S.; Lin, Y.H.; Zhang, B.P.; Wang, Y.; Nan, C.W. Controlled fabrication of BiFeO3 uniform microcrystals and their magnetic and photocatalytic behaviors. J. Phys. Chem. C 2010, 114, 2903–2908. [Google Scholar] [CrossRef]
  336. Xu, X.; Lin, Y.H.; Li, P.; Shu, L.; Nan, C.W. Synthesis and photocatalytic behaviors of high surface area BiFeO3 thin films. J. Am. Ceram. Soc. 2011, 94, 2296–2299. [Google Scholar] [CrossRef]
  337. Feng, Y.N.; Wang, H.C.; Shen, Y.; Lin, Y.H.; Nan, C.W. Magnetic and photocatalytic behaviors of Ba-doped BiFeO3 nanofibers. Int. J. Appl. Ceram. Technol. 2013, 11, 676–680. [Google Scholar] [CrossRef]
  338. Wang, H.C.; Lin, Y.H.; Feng, Y.N.; Shen, Y. Photocatalytic behaviors observed in Ba and Mn doped BiFeO3 nanofibers. J. Electroceram. 2013, 31, 271–274. [Google Scholar] [CrossRef]
Figure 1. Inorganic ABO3 perovksite structure.
Figure 1. Inorganic ABO3 perovksite structure.
Materials 13 05555 g001
Figure 2. A-site and X-site ions of hybrid organic-inorganic perovskites. Adapted from reference [15].
Figure 2. A-site and X-site ions of hybrid organic-inorganic perovskites. Adapted from reference [15].
Materials 13 05555 g002
Figure 3. Pulsed laser deposition (PLD) technique scheme Adapted from Ref. [84].
Figure 3. Pulsed laser deposition (PLD) technique scheme Adapted from Ref. [84].
Materials 13 05555 g003
Figure 4. Photograph of the PLD setup: the heated substrate and the material target.
Figure 4. Photograph of the PLD setup: the heated substrate and the material target.
Materials 13 05555 g004
Figure 5. Photograph of a TiO2 target.
Figure 5. Photograph of a TiO2 target.
Materials 13 05555 g005
Figure 6. The experimental setup of radiofrequency (RF)-assisted PLD (a) and RF plasma source (b) Adapted from reference [86].
Figure 6. The experimental setup of radiofrequency (RF)-assisted PLD (a) and RF plasma source (b) Adapted from reference [86].
Materials 13 05555 g006
Figure 7. Photographs of frozen material used as matrix-assisted pulsed laser evaporation (MAPLE) target.
Figure 7. Photographs of frozen material used as matrix-assisted pulsed laser evaporation (MAPLE) target.
Materials 13 05555 g007
Figure 8. The schematic setup of the MAPLE technique. Adapted from reference [86].
Figure 8. The schematic setup of the MAPLE technique. Adapted from reference [86].
Materials 13 05555 g008
Figure 9. Schematic illustration of (a) photocatalytic (PC-I. Light absorption; II. Electrons jump from the valence band (VB) to the conduction band (CB); III. Redox processes; IV. Bulk recombinations) and (b) photoelectrochemical (PEC-I. Light absorption; II. Electrons jump from VB to CB; III. Transfer of charge carriers to the electrodes surface; IV. Bulk recombinations; V. Redox processes) processes Adapted from Ref [163].
Figure 9. Schematic illustration of (a) photocatalytic (PC-I. Light absorption; II. Electrons jump from the valence band (VB) to the conduction band (CB); III. Redox processes; IV. Bulk recombinations) and (b) photoelectrochemical (PEC-I. Light absorption; II. Electrons jump from VB to CB; III. Transfer of charge carriers to the electrodes surface; IV. Bulk recombinations; V. Redox processes) processes Adapted from Ref [163].
Materials 13 05555 g009
Figure 10. Schematic diagram of the charge carriers formation in a photocatalytic process. Adapted from reference [165].
Figure 10. Schematic diagram of the charge carriers formation in a photocatalytic process. Adapted from reference [165].
Materials 13 05555 g010
Figure 11. Photoelectrochemical water-splitting systems with: (a) n-type semiconductor photoanode; (b) p-type semiconductor photocathode; (c) tandem system. Adapted from reference [163].
Figure 11. Photoelectrochemical water-splitting systems with: (a) n-type semiconductor photoanode; (b) p-type semiconductor photocathode; (c) tandem system. Adapted from reference [163].
Materials 13 05555 g011
Figure 12. The ideal VB and CB positions for water splitting reaction. Adapted from reference [168].
Figure 12. The ideal VB and CB positions for water splitting reaction. Adapted from reference [168].
Materials 13 05555 g012
Figure 13. J vs. V curves for a photoanode in dark and illumination in the presence/absence of a photocatalyst. Adapted from reference [175].
Figure 13. J vs. V curves for a photoanode in dark and illumination in the presence/absence of a photocatalyst. Adapted from reference [175].
Materials 13 05555 g013
Figure 14. The schematically representation of (a) photocatalyst; (b) photoelectrode loaded by co-catalysts. Adapted from reference [168].
Figure 14. The schematically representation of (a) photocatalyst; (b) photoelectrode loaded by co-catalysts. Adapted from reference [168].
Materials 13 05555 g014
Figure 15. The working principle of the reaction in the presence of photocatalysts (A) and photocurrent density vs. the applied potential curves of photoelectrodes (B) loaded with an oxidation co-catalyst for both dark and illuminated processes, where OC—oxidation cocatalyst, Euncat—activation energy without OC and Ecat—activation energy with OC. Adapted from references [168,228].
Figure 15. The working principle of the reaction in the presence of photocatalysts (A) and photocurrent density vs. the applied potential curves of photoelectrodes (B) loaded with an oxidation co-catalyst for both dark and illuminated processes, where OC—oxidation cocatalyst, Euncat—activation energy without OC and Ecat—activation energy with OC. Adapted from references [168,228].
Materials 13 05555 g015
Table 1. The influence of the synthesis method on the specific surface area (SSA) of different perovskite powders.
Table 1. The influence of the synthesis method on the specific surface area (SSA) of different perovskite powders.
CatalystPreparation MethodSSA
(m2/g)
Ref.
BaTiO3Solid state reaction0.4[44]
PbTiO3 0.5
LaAlO3Citrate4.0[55]
LaAl0.95Mn0.05O3 8.0
LaAl0.9Mn0.1O3 7.0
LaAl0.8Mn0.2O3 25.0
LaAl0.6Mn0.4O3 25.0
LaAl0.4Mn0.6O3 26.0
LaAl0.2Mn0.8O3 33.0
LaCrO3Solid state reaction1.9[45]
LaCrO3 andCitrate~5–7[47]
LaCr1-xMgxO3 with 0.1 ≤ x ≤ 0.5
LaCr0.5Mg0.5O3Citrate6.08[48]
LaCr0.5Mg0.5O3∙2MgO 13.6
LaCr0.5Mg0.5O3∙6MgO 24.2
LaCr0.5Mg0.5O3∙17MgO 36.9
LaNiO3Solid state reaction4.8[45]
LaNiO3Plasmochemical17.0[51]
La0.87Sr0.13Mn0.2Ni0.8O3-x Freeze drying12.7[52]
La0.4Sr0.6Mo0.1Ni0.9O3 (microstructured)Freeze drying10.3[53]
La0.4Sr0.6Mn0.4Ni0.6O3 (microstructured)Freeze drying15.9[53]
LaCuO3Solid state reaction0.6[45]
LaMn0.8Cu0.2O3Citrate19.0[50]
LaMn0.6Cu0.4O3Citrate14.0[50]
LaFe0.84Cu0.16O3Citrate4.0[49]
LaCo0.8Cu0.2O3Citrate21.0[38]
LaCo0.6Cu0.4O3Citrate13.0[38]
(Dy-Y)NiO3Mechanochemical9.5[51]
Table 2. The influence of the synthesis method on the specific surface area of different powder lanthanide cobaltate-based perovskites.
Table 2. The influence of the synthesis method on the specific surface area of different powder lanthanide cobaltate-based perovskites.
CatalystPreparation MethodSSA
(m2/g)
Ref.
LaCoO3Co-precipitation3.5[56]
LaCoO3Co-precipitation8.0[57]
LaCoO3Solid state reaction3.0[45]
LaCoO3Citrate11.3[58]
LaCoO3Citrate6.0[59]
LaCoO3Citrate15.0[50]
LaCoO3Combustion5.7[63]
LaCoO3Ultrasound spray combustion5.5[64]
LaCoO3Flame-pyrolysis43.0[61]
LaCoO3+xFlame-hydrolysis15.6–22.8[65]
La0.8Ca0.2CoO3Solid state reaction2.0[45]
La0.9Sr0.1CoO3Flame-pyrolysis52.0[61]
La0.8Sr0.2CoO3Solid state reaction4.7[45]
La0.8Sr0.2CoO3 aFreeze-drying16.5[53]
La0.66Sr0.34CoO3 a 17.4
La0.6Sr0.4CoO3Solid state reaction3.6[45]
La0.8Ba0.2CoO3 5.1
La0.53Sr0.47Fe0.2Co0.8O3 aFreeze-drying13.4[53]
La0.4Sr0.6Fe0.4Co0.6O3 a 10.4
La0.55Sr0.45Mn0.1Ni0.3Co0.6O3 a 15.0
La0.5Sr0.5Fe0.04Ni0.1Co0.86O3 a 19.6
La0.55Sr0.45Fe0.05Ni0.3Co0.65O3 a 15.2
La0.66Sr0.34Ni0.3Co0.7O3 a 18.8
La0.7Sr0.3Ni0.3Co0.7O3 a 22.7
La0.95Ce0.05CoO3Citrate8.7[58]
La0.9Ce0.1CoO3 10.0
La0.9Ce0.1CoO3Flame-pyrolysis62.0[61]
La0.8Ce0.2CoO3Solid state reaction3.1[45]
La0.8Ce0.2CoO3Citrate 14.2[58]
La0.7Ce0.3CoO3 14.3
La0.6Ce0.4CoO3 11.6
La0.5Ce0.5CoO3 18.0
PrCoO3Co-precipitation5.1[56]
NdCoO3 1.6
GdCoO3 2.1
a Microstructured powders.
Table 3. The influence of the synthesis method on the specific surface area of different lanthanide manganate-based perovskite powders.
Table 3. The influence of the synthesis method on the specific surface area of different lanthanide manganate-based perovskite powders.
CatalystPreparation MethodSSA
(m2/g)
Ref.
LaMnO3Co-precipitation~15.0[3]
LaMnO3+xCo-precipitation8.0[69]
LaMnO3Co-precipitation7.0[57]
LaMnO3Co-precipitation11.0[70]
LaMnO3Solid state reaction4.0[45]
LaMnO3Citrate5.6[49]
LaMnO3Citrate20.0[71]
LaMnO3Citrate20.0[72]
LaMnO3Citrate22.0[55]
LaMnO3Citrate68.0[66]
LaMnO3Citrate22.0[50]
LaMnO3+xFlame-hydrolysis15.6–22.8[65]
LaMnO3Flame-pyrolysis56.0[61]
LaMnO3Ultrasound spray combustion21.8[64]
La0.8Ca0.2MnO3Solid state reaction6.7[45]
La0.9Sr0.1MnO3Flame-pyrolysis51.0[61]
La0.8Sr0.2MnO3Flame-pyrolysis70.0[61]
La0.8Sr0.2MnO3Solid state reaction8.6[45]
La0.8Sr0.2MnO3+xCo-precipitation5.0[73]
La0.6Sr0.4MnO3Citrate18.7[71]
La0.6Sr0.4MnO3Solid state reaction3.3[45]
La0.6Sr0.4MnO3 aCitrate4.32[74]
SrTi0.8Zr0.1Mn0.1O3Co-precipitation15.0[75]
La(Mn,Pd)O3 (2.9 wt.% Pd)Citrate method12.0[67]
La(Mn,Pd)O3 (2.32 wt.% Pd)Flame pyrolysis32.0[67]
La(Mn,Pd)O3 (2.37 wt.% Pd)Solution combustion1.0[67]
La(Mn,Pd)O3 (2.11 wt.% Pd)Ultrasonic spray combustion39.0[67]
La0.9Ce0.1MnO3Citrate32.0[71]
La0.9Ce0.1MnO3Flame-pyrolysis84.0[61]
(La-Ce)MnO3Mechanochemical19.0[51]
PrMnO3+xCo-precipitation2.5[69]
NdMnO3+xCo-precipitation2.4[69]
NdMnO3Citrate20.0[72]
SmMnO3 19.0
Sm0.9Sr0.1MnO3 20.0
Sm0.7Sr0.3MnO3 14.0
Sm0.5Sr0.5MnO3 13.0
La0.9Eu0.1MnO3Citrate26.4[71]
GdMnO3+xCo-precipitation5.3[69]
(Dy-Y)MnO3Mechanochemical14.0[51]
a 1D non-porous.
Table 4. The influence of the synthesis method on the specific surface area of lanthanide ferrite-based perovskite powders.
Table 4. The influence of the synthesis method on the specific surface area of lanthanide ferrite-based perovskite powders.
CatalystPreparation MethodSSA
(m2/g)
Ref.
LaFeO3Co-precipitation10.0[57]
LaFeO3Co-precipitation5.5[3]
LaFeO3+xCo-precipitation3.5[69]
LaFeO3Co-precipitation20.9[76]
LaFeO3Solid state reaction3.1[45]
LaFeO3Citrate2.9[77]
LaFeO3Citrate3.0[78]
LaFeO3Citrate19.5[76]
LaFeO3Citrate21.0[79]
LaFeO3Combustion3.0[63]
LaFeO3+xFlame-hydrolysis15.6[65]
LaFeO3Flame-pyrolysis38.0[61]
LaFeO3Ultrasound spray combustion9.8[64]
LaFe0.9Mg0.1O3Citrate4.3[77]
LaFe0.8Mg0.2O3 5.5
LaFe0.7Mg0.3O3 7.9
LaFe0.6Mg0.4O3 9.7
LaFe0.5Mg0.5O3 5.3
La0.9Ca0.1FeO3Co-precipitation14.0[76]
La0.9Ca0.1FeO3Citrate6.0[78]
La0.9Ca0.1FeO3Citrate17.8[76]
La0.9Ca0.1FeO3Citrate38.0[79]
La0.8Ca0.2FeO3Co-precipitation8.3[76]
La0.8Ca0.2FeO3Citrate5.0[78]
La0.8Ca0.2FeO3Citrate38.6[76]
La0.8Ca0.2FeO3Citrate38.0[79]
La0.7Ca0.3FeO3Co-precipitation5.4[76]
La0.7Ca0.3FeO3Citrate3.0[78]
La0.7Ca0.3FeO3Citrate38.6[76]
La0.7Ca0.3FeO3Citrate38.0[79]
La0.6Ca0.4FeO3Co-precipitation8.3[76]
La0.6Ca0.4FeO3Citrate31.2[76]
La0.6Ca0.4FeO3Citrate5.0[78]
La0.6Ca0.4FeO3Citrate33.0[79]
La0.5Ca0.5FeO3Citrate0.7[78]
La0.8Sr0.2FeO3Solid state reaction4.7[45]
La(Fe,Pd)O3 (2.4 wt.% Pd)Citrate14.0[67]
La(Fe,Pd)O3 (2.28 wt.% Pd)Flame pyrolysis22.0
La(Fe,Pd)O3 (1.25 wt.% Pd)Ultrasonic spray combustion27.0
La(Fe,Pd)O3 (2.47 wt.% Pd)Solution combustion1.6
(La-Ce)FeO3Mechanochemical5.3[51]
PrFeO3+xCo-precipitation5.8[69]
NdFeO3+xCo-precipitation4.5[69]
NdFeO3Citrate2.3[77]
SmFeO3Citrate4.3[77]
GdFeO3+xCo-precipitation5.6[69]
(Dy-Y)FeO3Mechanochemical8.6[51]
Table 5. The effect of the preparation methods on the SSA of supported perovskites.
Table 5. The effect of the preparation methods on the SSA of supported perovskites.
CatalystSupportMorphology of the SupportPreparation MethodSSA
(m2/g)
Ref.
LaMnO3foil Fe2Cr20Al5monolithWet impregnation23.3[80]
LaMnO3ZrO2powderSolution combustion method132.5[82]
30% LaMnO3(5%La2O3/Al2O3)powderDeposition precipitation88.0[70]
20% LaMnO3MgO 25.0[70]
La0.8Sr0.2MnO3+xMgAl2O4powderWet impregnation34.0[73]
La0.8Sr0.2MnO3+xNiAl2O4 22.0
La0.8Sr0.2MnO3+xCoAl2O4 18.0
La0.95Ag0.05MnO3foil Fe2Cr20Al5monolithWet impregnation27.4[80]
La0.9Ag0.1MnO3 30.9
La0.8Ag0.2MnO3 29.4
La0.7Ag0.3MnO3 31.5
LaFeO3FeCr(20%)Al(5%)monolithCitrate method7.7[81]
La0.66Sr0.34Ni0.29Co0.69Mn0.02O3(20%)(47% Al2O3–52% SiO2)fiberFreeze-drying18.0[83]
La0.66Sr0.34Ni0.29Co0.69Mn0.02O3(15%)(47% Al2O3–52% SiO2) 23.0
La0.66Sr0.34Ni0.29Co0.69Fe0.02O3 (27%)(95% Al2O3–5% SiO2) 27.0
La0.66Sr0.34Ni0.29Co0.69Fe0.02O3 (14%)(95% Al2O3–5% SiO2) 22.0
La0.66Sr0.34Ni0.29Co0.69Fe0.02O3(15%)(47% Al2O3–52% SiO2) 24.0
La0.66Sr0.34Ni0.29Co0.69Fe0.02O3(12%)(47% Al2O3–52% SiO2) 27.0
Table 6. Experimental conditions for different types of perovskite materials grown by PLD and PLD-RF techniques and the thickness of the films obtained.
Table 6. Experimental conditions for different types of perovskite materials grown by PLD and PLD-RF techniques and the thickness of the films obtained.
PerovskiteSupportλ (nm)/ν (Hz)Fluence
(J/cm2)
PO2 a
(mBar)
T b
(°C)
Film Thickness
(nm)
Ref.
BaTiO3MgO (001)248/102–44 × 10−31000115[88,89]
BaTiO3SrTiO3 (001)248/102–43 × 10−1800115[88,89]
BaTiO3SrTiO3 (001)248/102–42 × 10−3800220[88,89]
SrTiO3LaAlO3 (100)248/51.31.7 × 10−4660-[90]
BaZrO3Si193/1–35.20.26051–53[91]
BaZrO3SrTiO3 (100)193/1–35.20.26501–53[91]
Y1-xSrxMnO3
(x = 1, 0.9, 0.8, 0.7)
SrTiO3 (100)248/320.01–0.138000.6–0.65[92]
Y1-xSrxMnO3
(x = 1, 0.9, 0.8, 0.7)
LaAlO3 (100)248/320.01–0.138000.6–0.65[92]
Y1-xSrxMnO3
(x = 1, 0.9, 0.8, 0.7)
NdGaO3 (101)248/320.01–0.138000.6–0.65[92]
La0.67Ca0.33MnO3LaAlO3-/5n.s. d0.26500–700100[93]
La0.67Ca0.33MnO3 RF cLaAlO3 (100)n.s.n.s.Pressure of 0.053 (60:40 = Ar:O2 or pure Ar)850n.s.[93]
La0.67Ca0.33MnO3 RF cNdGaO3 (110)n.s.n.s.Presure of 0.053 (60:40 = Ar:O2 or pure Ar)850n.s.[93]
La0.67Ca0.33MnO3NdGaO3248/80.2–0.250.13600–80085[94]
La0.67Ca0.33MnO3 RF cSrTiO3 (100)n.s.n.s.Presure of 0.053 (60:40 = Ar:O2 or pure Ar)850n.s.[93]
SrFeO3SrTiO3 (111)248/22.30.13700n.s.[95]
SrTi1−xFexO3−y
(x = 0.2–0.5)
Sapphire (Al2O3)248/81.50.13700200–300[96]
LaNiO3SrTiO3 (100)248/520.4825100[97]
LaNiO3LaAlO3 (100)248/51.30.35660n.s.[90]
Y(Ni0.5Mn0.5)O3SrTiO3 (001)248/21.50.6550–85070[98]
Y(Ni0.5Mn0.5)O3SrTiO3 (110)248/2–2020.6550–85070[98]
Y(Ni0.5Mn0.5)O3SrTiO3 (111)248/2–2020.6550–85070[98]
Ca0.25Cu0.75TiO3SrTiO3 (100)248/1030.16600–800250[99]
SrRuO3SrTiO3 (100)193/1–35.20.2590–650n.s.[91]
SrRuO3SrTiO3 (111)2482.30.13700n.s.[95]
PbTiO3Si (100)248/580.13–0.2530530[100]
PbZr0.2Ti0.8 O3SrRuO3/SrTiO3 (100)193/1–35.20.2587n.s.[91]
Pb(Zr0.45Ti0.55)O3Pb(Zr0.45Ti0.55)O3/Pt/Ti/SiO2/Si prepared by citrate method248/101.21.3 × 10−4RT e600[101]
BiFeO3Pt/TiO2/SiO2/Si355/2.5n.s.0.07450300[102,103]
BiFeO3SrTiO3 (001)355/2.5n.s.0.0158070[102,103]
BiFeO3Pt(111)/TiO2/SiO2/Si(100)248/52.5n.s.450230[104]
BiFeO3SrTiO3 (100)248/520.07750400[105]
BiFeO3SrRuO3/SrTiO3248n.s.0.07550–800280–300[106]
BiFeO3Nb-doped SrTiO3 (100)n.s.n.s.0.14650106.5[107,108]
BiFeO3SrTiO3 (100)n.s.n.s.0.1465054.3[107,108]
BiFeO3DyScO3 (110)n.s.n.s.0.1465034.1[107,108]
BiFeO3Nb-doped SrTiO3 (100)248/5n.s.0.04–0.467030–80[109]
BiFeO3Pt/TiO2/SiO2/Si248/102.5n.s.RT e150[110]
BiFeO3Nb-doped SrTiO3 (100)248n.s.0.0250010[111]
BiFeO3SrRuO3/SrTiO3 (001)248n.s.0.01–0.26650–750n.s.[112]
BiFeO3Si (100)248/5n.s.n.s.67050–100[113]
BiFeO3Pt/TiO2/SiO2/Si248/31.50.53625100[114,115]
BiFeO3SrRuO3/SrTiO3 (111)248/31.50.53625100[114,115]
BiFeO3Pt coated Al2O3248/31.50.53625100[114,115]
(1-x)BiFeO3-xPbTiO3PbTiO3/Pt/Ti/SiO2/Si248/52.50.2550230[116]
(1-x)BiFeO3-xPbTiO3
(x = 0.7, 0.8, 0.9)
Pt/Ti/SiO2248/560.2545400[117]
LaCo1-xCrxO3LaAlO3 (100)248/15n.s.0.083650450–530[118]
LaMnO3SrTiO3 (001)n.s.n.s.0.01370030[119]
LaMnO3LaAlO3 (001)248/2n.s.0.0170014–90[120]
La0.8Ca0.2MnO3LaAlO3 (100)248/53.20.4800200[121]
La0.7Sr0.3MnO3Si (100)248/10n.s.0.27650200[122]
La0.67Sr0.33MnO3LaAlO3308/520.27500–800130[123]
La0.67Sr0.33MnO3SrTiO3308/520.27500–800130[123]
LaFeO3SrTiO3 (100)248/41.90.467025–35[124]
LaFeO3GdScO3 (110)24810.13700100[125]
LaFeO3SrTiO3 (100)248/42.40.267054[126]
LaFeO3SrTiO3 (100)248/55.50.07800400[127]
LaFeO3Nb-doped SrTiO3193/52.20.05–0.9750n.s.[128]
LaFeO3SrTiO3 (100)248/100.24 × 10−570065[129]
LaFeO3LaAlO3 (100)248/100.24 × 10−570065[129]
La1-xSrxCoO3Si (001)266/1020.05740120[130]
La1-xSrxCoO3MgO (001)266/1020.05740120[130]
La1-xSrxCoO3
(x = 0, 0.1, 0.2)
Si (100)248/1020.13600n.s.[131]
La1-xSrxCoO3
(x = 0, 0.1, 0.2)
Si (100)266/1020.05660100[132]
Sr1-xLaxRu1-xFexO3
(x = 0.05, 0.1, 0.2, 0.3)
SrTiO3 (100)248/42.50.2–0.3375060[133]
Bi0.9La0.1Fe0.95Mn0.05O3Pt (111)/Ti/SiO2/Si2481.57 × 10−3450–650250[134]
(Bi1-xLax)(Fe1-xAlx)O3 (x = 0, 0.1, 0.2, 0.3, 0.4)Nb-doped SrTiO3 (001)266n.s.0.065600n.s.[135]
La0.8Ce0.2MnO3LaAlO3 (001)248/820.4800150[136]
Bi1-xPrxFeO3
(x = 0, 0.05, 0.1, 0.15)
Pt/SiO2355/52.5n.s.450200[137]
NdNiO3MgO (100)248/101.50.15675500[138]
NdNiO3SrTiO3 (100)248/101.50.15675500[138]
NdNiO3NdGaO3 (110)248/101.50.15675500[138]
NdNiO3NdGaO3 (001)-/101.9390030–50[139]
Bi0.9Sm0.1Fe0.95Co0.05O3Pt/TiO2/SiO2/Si248/102–50.13700–750300–360[140]
Bi0.9Sm0.1Fe0.95Co0.05O3Pt/TiO2/SiO2/Si248/102–50.13700–750300–360[140]
Bi1-xSmxFeO3 (x = 0.05, 0.1, 0.12, 0.14, 0.16)Pt (111)/SiO2355/5n.s.0.04450200[141]
TmMnO3SrTiO3 (110)248n.s.0.194020[142]
a PO2—the oxygen pressure during the deposition process; b T—the substrate temperature during the deposition process; c RF—prepared via PLD-RF; d n.s.—not specified; e RT—room temperature.
Table 7. Perovskites prepared by MAPLE.
Table 7. Perovskites prepared by MAPLE.
MaterialTarget ConcentrationSubstrateλ (nm)/
ν (Hz)
Fluence (J/cm2)PN2 (mbar)DT-S (cm)Ref.
rGO/BiFeO33 wt.% BiFeO3
5 wt.% GO
F-doped SnO2266/100.40.2 4[148]
rGO/LaFeO33 wt.% LaFeO3
5 wt.% GO
F-doped SnO2266/100.40.2 4[148]
CH3NH3PbI3PbI2:MAI = 1:3ITO IR/20.125–0.1351 × 10−37[149]
CH3NH3PbI3PbI2:MAI = 1:1 in DMSO and MEGFTO/NiOx ----[150]
Abbreviations: DT-S—the distance between the material target and the substrate; rGO—reduced graphene oxide; MAI—methylammonium iodide; ITO—indium tin oxide; IR—infrared radiation; DMSO—dimethyl sulfoxide; MEG—ethylene glycol; FTO—fluorine-doped tin oxide.
Table 8. Photoelectrodes based on perovskites for photoelectrochemical water-splitting reaction.
Table 8. Photoelectrodes based on perovskites for photoelectrochemical water-splitting reaction.
MaterialElectrode TypeEg a (eV)Electrolyte/Light Source/Intensity (mW/cm2)PerformanceStabilityRef.
BaTiO3Photo-anode3.110.1 M NaOH (pH = 13)/Xe arc UV-Vis lamp/180 0.5 VSCE: Jph ≈ 0.07 mA/cm2; n.s.b[188]
PbTiO3-TiO2Photo-anode 2.78–3.60.1 M KOH/Xe lamp Vis light 1.23 VRHE: Jph ≈ 0.3 mA/cm2 IPCEc ≈ 70% at 380 nm
IPCE ≈ 38% at 420 nm
IPCE < 1% at 500 nm
Onset potential 0.3 VRHE
Stable after 300 s[189]
SrTiO3 nanocubesPhoto-anode3.430.1 Na2SO4 (pH = 7)/AM1.5/1000 VAg/AgCl: Jph ≈ 0.5 uA/cm2
0.9 VAg/AgCl: Jph ≈ 4 uA/cm2;
60 μmol/h O2
n.s.[190]
SrTiO3Photo-anode3.40.1 M Na2SO4/AM1.5/1000 VAg/AgCl: Jph ≈ 50 μA/cm2
1.5 VAg/AgCl: Jph ≈ 0.5 mA/cm2 IPCE ≈ 10% at 350 nm
IPCE < 1% for λ > 400 nm
n.s.[191]
SrTiO3—carbon quantum dotsPhoto-anoden.s.0.1 M Na2SO4/AM1.5/1000 VAg/AgCl: Jph ≈ 110 μA/cm2
1.5 VAg/AgCl: Jph ≈ 1.7 mA/cm2
IPCE ≈ 14% at 350 nm
IPCE ≈ 1% at 860 nm
n.s.[192]
BiFeO3Photo-anode2.10.2 M Na2SO4/sunlight 300 W Xenon lamp0.8 VAg/AgCl: Jph ≈ 0.12 μA/cm2Stable after 400 s[201]
H2 treated BiFeO3Photo-anode2.00.2 M Na2SO4/sunlight, 300 W Xenon lamp0.8 VAg/AgCl: Jph ≈ 0.69 μA/cm2 Stable after 400 s[201]
BiFeO3Photo-anode2.40.2 M Na2SO4 (pH = 6.5)/AM1.5/1001 VAg/AgCl: Jph≈ 0.17 mA/cm2
IPCE≈ 17% at 420 nm
Onset potential 0.1 VAg/AgCl
n.s.[202]
Ni-B/BiFeO3Photo-anoden.s.0.1 M potassium borate (pH = 9.2)1 VAg/AgCl: Jph ≈ 0.72 mA/cm2Stable after 3 h[202]
BFO/SrRuO3Photo-anode2.741M Na2SO4/250 mW/cm20.64 VAg/AgCl: Jph ≈ 10 μA/cm2
Onset potential 0.18 VAg/AgCl
n.s.[221]
Ti-doped BiFeO3Photo-anode1.971 M NaOH/300W UV Xe lamp1.23 VRHE: Jph ≈ 0.04mA/cm2
Onset potential 0.81 VRHE
Stable after 3600 s[208]
Y-doped BiFeO3Photo-anoden.s.0.5 M NaOH (pH = 13)/laser diode 405 nm (5 mW)1.4 VRHE: Jph = 0.72 mA/cm2Stable after 900 s[209]
BiFeO3/TiO2/FTOPhoto-anoden.s.1M NaOH/300 W Xenon lamp Vis light1.5 VSCE: Jph ≈ 15 mA/cm2
Onset potential ≈ 0.6 VSCE
Stable after 300 s[222]
BiFeO3/TiO2/FTOPhoto-anoden.s.1M NaOH/AM1.5/1001.5 VSCE: Jph ≈ 17 mA/cm2
Onset potential ≈ 0.6 VSCE
Stable after 300 s[222]
WO3/BiVO4/BiFeO3Photo-anoden.s.0.5M Na2SO4/AM1.5/1002.53 VRHE: Jph = 46.9 mA/cm2Stable after 200 s[212]
LaFeO3Photo-cathode2.160.1 M NaOH (pH = 13)/AM1.5/1000.73 VRHE: Jph = −0.1 mA/cm2Stable after 1 h[223]
LaFeO3Photo-cathode2.160.1 M NaOH (pH = 13)/AM1.5/1000.5 VRHE: Jph ≈ −0.2 mA/cm2Stable after 1 h[223]
LaFeO3Photo-anode2.080.1 M KOH/500 W Xenon lamp, Vis light/1001.1 VAg/AgCl: Jph≈ 1.2 mA/cm2 Onset potential 0.48 VAg/AgClStable after 330 s[220]
LaFeO3Photo-cathode2.40.1 M NaOH 0.26 VRHE: Jph ≈ 0.16 mA/cm2 Onset potential 1.2 VRHEStable after 21 h[224]
LaFeO3Photo-cathode1.951M Na2SO4/AM1.5/1001.7 VAg/AgCl:
Jph ≈ 8.2 mA/cm2
decreases to 50% after 30 min[225]
LaFeO3Photo-anode2.070.1 M NaOH (pH = 13)/laser diode 405 nm (5mW)1 VAg/AgCl: Jph = 1.6 mA/cm2n.s.[128]
LaFe0.9Mn0.1O3Photo-anode~2.080.1 M KOH/500 W Xenon lamp, Vis light/1001.1VAg/AgCl: Jph ≈ 1.5 mA/cm2
Onset potential 0.34 V
Stable after 330 s[220]
LaFe0.9Co0.1O3Photo-anode~2.080.1 M KOH/500 W Xenon lamp, Vis light/1001.1 VAg/AgCl: Jph ≈ 1.8 mA/cm2
Onset potential 0.27 V
Stable after 330 s[220]
LaFe0.9Cu0.1O3Photo-anode~2.080.1 M KOH/500 W Xenon lamp, Vis light/1001.1 VAg/AgCl: Jph ≈ 2 mA/cm2
Onset potential 0.27 V
Stable after 330 s[220]
p-LaFeO3/n-Fe2O3Photo-cathode/
Photo-anode
n.s.1M NaOH/AM1.5/1000 VRHE: Jph = 64.5 μA/cm2
11.5 μmol/h H2
5.7 μmol/h O2
Stable after 120 h[218]
LaFeO3Photo-cathode2.40.1 M NaOH/AM1.5/1000.6 VRHE: Jph ≈ −0.04 mA/cm2n.s.[226]
Ag-LaFeO3Photo-cathoden.s.1 M NaOH/AM1.5/1000.6 VRHE: Jph ≈ −0.074 mA/cm2n.s.[226]
LaFeO3Photo-cathode2.070.1 M Na2SO4/AM1.5/1000.6 VRHE: Jph ≈ −4.78 μA/cm2decreases to 88.6% after 2750 s[227]
FTO/Au/LaFeO3Photo-cathoden.s.0.1 M Na2SO4/AM1.5/1000.6 VRHE: Jph ≈ −19.60 μA/cm2decreases to 91% after 2750 s [227]
a Eg—band gap; b n.s.—not specified; c ICPE—Incident Photon to Current Efficiency.
Table 9. Methane combustion activity of undoped perovskite-type materials expressed as the temperature where 50% conversion is reached (T50).
Table 9. Methane combustion activity of undoped perovskite-type materials expressed as the temperature where 50% conversion is reached (T50).
CatalystSSA (m2/g)Reaction ConditionsT50
(°C)
Ea a (kJ/mol)Ref.
LaCoO33.02 vol. % CH4 in air, 45,000–50,000/h52522.1[45]
LaCoO33.51 vol. % CH4, 4 vol. % O2 in He, 135,000/h709~104 [56]
LaCoO35.73 vol. % CH4, 7.2 vol. % O2 in N2, 113 cm3/min, 0.5 g catalyst~647n.s. b[63]
LaCoO381 vol. % CH4 in air 545n.s.[57]
LaCoO315.00.4 vol. % CH4, 10 vol. % O2 in N2, 40,000 Ncm3/h x gcat~567n.s.[50]
LaCoO3+x15.6–22.810 cm3 (1.04 vol. % CH4 in He) with 10 cm3 of air, 0.2 g catalyst466n.s.[65]
LaCoO311.31 vol. % CH4 in air, 45,000 mL/(h gcat), 0.1 g catalyst60097[58]
LaCoO36.01 vol. % CH4, 4 vol.% O2 in He, 20,000–200,000/h, 0.1 g catalyst~600n.s.[59]
LaCoO3430.34 vol.% CH4, 33.3 vol.% air in He, 30 Ncm3/min, 0.15 g catalyst449n.s.[61]
LaCoO35.51 vol.% CH4, 4 vol.% O2 in He, 14,150/h, 0.1 g catalyst560n.s.[64]
PrCoO35.11 vol.% CH4, 4 vol.% O2 in He, 135,000/h903~110[56]
NdCoO31.6658~103
GdCoO32.1733~99
LaMnO34.02 vol.% CH4 in air, 45,000–5000/h57921.8[45]
LaMnO3~151.5 vol.% CH4, 4.2 vol.% O2 in He), 200 cm3/min, 0.004 g catalystn.s.73[3]
LaMnO35.63.2 vol.% CH4, 12.8 vol.% O2 in Ar, 73.5 mL/min 45792 [49]
LaMnO3+x8.01 vol.% CH4, 4 vol.% O2 in He, 135,000/h682~82[69]
LaMnO371 vol.% CH4 in air 580n.s.[57]
LaMnO3n.s.1 vol.% CH4 in air, ~50,000 cm3/(h gcat)~577n.s.[75]
LaMnO320.00.5 vol.% CH4, 10 vol.% air in N2, 40 Ncm3/min, 0.2 g catalyst~440n.s.[71]
LaMnO322.00.4 vol.% CH4, 10 vol.% O2 in N2, 40,000 Ncm3/(h gcat)~506n.s.[50]
LaMnO320.00.4 vol.% CH4, 10 vol.% O2 in N2~50723.3[72]
LaMnO311.00.4 vol.% CH4, 10 vol.% O2 in N257524.4[70]
LaMnO3+x15.6–22.810 cm3 (1.04 vol.% CH4 in He) with 10 cm3 of air, 0.2g catalyst489n.s.[65]
LaMnO3n.s.0.4 vol.% CH4, 10 vol.% O2 in N2, 40,000 Ncm3/(h gcat), 0.4 g catalyst51197.5[238]
LaMnO322.00.4 vol.% CH4, 10 vol.% O2 in N2, 40,000 cm3/(h gcat), 0.4 g catalyst~50723.3[55]
LaMnO3560.34 vol.% CH4, 33.3 vol.% air in He, 30 Ncm3/min, 0.15 g catalyst435n.s.[61]
LaMnO321.81 vol.% CH4, 4 vol.% O2 in He, 14,150/h, 0.1 g catalyst515n.s.[64]
LaMnO3681 vol.% CH4, 4 vol.% O2 in N2, 40,000 cm3/(h gcat), 0.15 g catalyst446n.s.[66]
30% LaMnO3/
(5% La2O3/Al2O3)
88.00.4 vol.% CH4, 10 vol.% O2 in N253218.2[70]
20% LaMnO3/MgO25.053323.3
LaMnO3-ZrO2132.52 vol.% CH4, 16 vol.% O2 in He, 50 Ncm3/min, 0.1 g catalyst595n.s.[82]
Pd/LaMnO3-ZrO274.6570n.s.
0.5% Pt/LaMnO3630.34 vol.% CH4, 33.3 vol.% air in He, 30 Ncm3/min, 0.15 g catalyst426n.s.[61]
0.5% Pd/LaMnO353445n.s.
LaMnO3/foil Fe2Cr20Al523.31 vol.% CH4 in air,
64,410 cm3/(h gcat), 25.7 g catalyst
566n.s.[80]
PrMnO3+x2.51 vol.% CH4, 4 vol.% O2 in He, 135,000/h711~89 [69]
NdMnO3+x2.4695~83
NdMnO320.00.4 vol.% CH4, 10 vol.% O2 in N2~58719.3[72]
GdMnO3+x5.31 vol.% CH4, 4 vol.% O2 in He, 135,000/h677~79[69]
SmMnO319.00.4 vol.% CH4, 10 vol.% O2 in N2~52717.1[72]
LaFeO33.12 vol.% CH4 in air, 45,000–5000/h57118.2[45]
LaFeO35.51.5 vol.% CH4, 4.2 vol.% O2 in He, 200 cm3/min, 0.004g catalystn.s.75[3]
LaFeO33.0n.s.~672n.s.[63]
LaFeO3101 vol.% CH4 in air 545n.s.[57]
LaFeO32.90.4 vol.% CH4, 10 vol.% O2 in N2, 40,000 Ncm3/(h gcat)52920.76[77]
LaFeO33.00.4 vol.% CH4, 10 vol.% O2 in N2, 40,000 Ncm3/(h gcat), 0.4 g catalyst52921.07[78]
LaFeO320.9n.s.608n.s.[76]
LaFeO319.5 508
LaFeO3380.34 vol.% CH4, 33.3 vol.% air in He, 30 Ncm3/min, 0.15 g catalyst495n.s.[61]
LaFeO39.81 vol.% CH4, 4 vol.% O2 in He, 14,150/h, 0.1 g catalyst625n.s.[64]
LaFeO321.037,000 ppmv CH4, 23.22 vol.% O2 in He512105.7[79]
LaFeO3+x15.610 cm3 (1.04 vol.% CH4 in He) with 10 cm3 of air, 0.2g catalyst472n.s.[65]
LaFeO3+x3.51 vol.% CH4, 4 vol.% O2 in He, 135,000/h678~105[69]
LaFeO3/FeCr(20%)Al(5%)7.71 vol.% CH4 in air, 5800/h 577101.8[81]
PrFeO3+x5.81 vol.% CH4, 4 vol.% O2 in He, 135,000/h71786[69]
NdFeO3+x4.5718~109
NdFeO32.30.4 vol.% CH4, 10 vol.% O2 in N2, 40,000 Ncm3/(h gcat)55620.76[77]
GdFeO3+x5.61 vol.% CH4, 4 vol.% O2 in He, 135,000/h707~89[69]
SmFeO34.30.4 vol.% CH4, 10 vol.% O2 in N2, 40,000 Ncm3/(h gcat)55823.66[77]
LaCuO30.62 vol.% CH4 in air, 45,000–5000/h67223.8[45]
LaNiO34.870219.4
LaNiO3n.s.1.5 vol.% CH4, 4.2 vol.% O2 in He, 200 cm3/min, 0.004 g catalystn.s.79[3]
LaNiO317.00.4 vol.% CH4, 2 vol.% O2 in He, 60,000/h, 1.5 g catalyst~60018.7 [51]
LaCrO31.92 vol.% CH4 in air, 45,000–5000/h78028.8[45]
LaCrO3n.s.1.5 vol.% CH4, 4.2 vol.% O2 in He, 200 cm3/min, 0.004g catalystn.s.142 [3]
LaCrO3~5–71.5 vol.%, 18 vol.% in He, 1.2cm3/s692n.s.[47]
LaRuO3n.s.1.5 vol.% CH4, 4.2 vol.% O2 in He, 200 cm3/min, 0.004g catalyst-95[3]
LaAlO34.00.4 vol.% CH4, 10 vol.% O2 in N2, 40,000 cm3/(h gcat), 0.4 g catalyst~65228.1[55]
BaCeO3-ZrO245.62 vol.% CH4, 16 vol.% O2 in He, 50 Ncm3/min, 0.1 g catalyst490n.s.[82]
Pd/BaCeO3-ZrO226.4512n.s.
BaTiO30.45 vol.% CH4 in air, 16,000/h74485.8[44]
PbTiO30.5697104.9
a Ea = activation energy; b n.s.—not specified.
Table 10. Performances of A-site-doped perovskite-type materials for the catalytic combustion of methane.
Table 10. Performances of A-site-doped perovskite-type materials for the catalytic combustion of methane.
CatalystSSA (m2/g)Reaction Conditions T50
(°C)
Ea (kJ/mol)Ref.
La0.9Sr0.1CoO3520.34 vol.% CH4, 33.3 vol.% air in He, 30 Ncm3/min, 0.15 g catalyst454n.s. b[61]
La0.8Sr0.2CoO34.72 vol.% CH4 in air, 45,000–5000/h51821.3[45]
La0.8Sr0.2CoO316.54 vol.% CH4 in air, 4.2–5 cm3/s,
1g catalyst
640 an.s.[53]
La0.75Sr0.25CoO3n.s.1.5 vol.% CH4, 4.2 vol.% O2 in He, 200 cm3/min, 0.004 g catalystn.s.81[3]
La0.66Sr0.34CoO317.44 vol% CH4 in air, 4.2–5 cm3/s,
1 g catalyst
675 an.s.[53]
La0.6Sr0.4CoO33.62 vol.% CH4 in air, 45,000–5000/h57019.0[45]
La0.5Sr0.5CoO3n.s.1.5 vol.% CH4, 4.2 vol.% O2 in He), 200 cm3/min, 0.004g catalystn.s.70[3]
La0.8Ba0.2CoO35.12 vol.% CH4 in air, 45,000–5000/h53516.9[45]
La0.8Ca0.2CoO32.060618.1
La0.95Ce0.05CoO38.71 vol.% CH4 in air,
45,000 mL/(h gcat), 0.1 g catalyst
53286[58]
La0.9Ce0.1CoO310.051583
La0.9Ce0.1CoO3620.34 vol.% CH4, 33.3 vol.% air in He, 30 Ncm3/min, 0.15 g catalyst447n.s.[61]
La0.8Ce0.2CoO33.12 vol.% CH4 in air, 45,000–5000/h49919.7[45]
La0.8Ce0.2CoO314.21 vol.% CH4 in air,
45,000 mL/(h gcat), 0.1 g catalyst
52097[58]
La0.7Ce0.3CoO314.350581
La0.6Ce0.4CoO311.6530n.s.
La0.9Sr0.1MnO3510.34 vol.% CH4, 33.3 vol.% air in He, 30 Ncm3/min, 0.15 g catalyst419n.s.[61]
La0.8Sr0.2MnO38.62 vol.% CH4 in air, 45,000–5000/h51019.7[45]
La0.8Sr0.2MnO3700.34 vol.% CH4, 33.3 vol.% air in He, 30 Ncm3/min, 0.15 g catalyst434n.s.[61]
La0.8Sr0.2MnO3+x5.01 vol.% CH4, 4 vol.% O2 in He, 135,000/h624104[73]
La0.8Sr0.2MnO3+x/MgAl2O434.061992
La0.8Sr0.2MnO3+x/NiAl2O422.064298
La0.8Sr0.2MnO3+x/CoAl2O418.0707114
La0.75Sr0.25MnO3n.s.1.5 vol.% CH4, 4.2 vol.% O2 in He), 200 cm3/min, 0.004 g catalystn.s.65[3]
La0.6Sr0.4MnO33.32 vol.% CH4 in air, 45,000–5000/h48220.1[45]
La0.6Sr0.4MnO318.70.5 vol.% CH4, 10 vol.% air in N2, 40 Ncm3/min, 0.2 g catalyst~470n.s.[71]
La0.6Sr0.4MnO34.325 vol.% CH4, 30 vol.% O2 in N2, 50,000 cm3/(h gcat), 0.05 g catalyst480136[74]
La0.6Sr0.4MnO333.5385102
La0.6Sr0.4MnO348.936067.3
La0.5Sr0.5MnO3n.s.1.5 vol.% CH4, 4.2 vol.% O2 in He), 200 cm3/min, 0.004 g catalystn.s.60[3]
La0.8Ca0.2MnO36.72 vol.% CH4 in air, 45,000–5000/h54318.9[45]
La0.9Ce0.1MnO3320.5 vol.% CH4, 10 vol.% air in N2, 40 Ncm3/min, 0.2 g catalyst~440n.s.[71]
La0.9Ce0.1MnO3840.34 vol.% CH4, 33.3 vol.% air in He, 30 Ncm3/min, 0.15 g catalyst433n.s.[61]
(La-Ce)MnO319.00.4 vol.% CH4, 2 vol.% O2 in He, 60,000/h, 1.5g catalyst~73021 [51]
La0.9Eu0.1MnO326.40.5 vol.% CH4, 10 vol.% air in N2, 40 Ncm3/min, 0.2 g catalyst~425n.s.[71]
Sm0.9Sr0.1MnO320.00.4 vol.% CH4, 10 vol.% O2 in N2~55720.8[72]
Sm0.7Sr0.3MnO314.0 ~52718.6
(Dy-Y)MnO314.00.4 vol.% CH4, 2 vol.% O2 in He, 60,000/h, 1.5g catalyst~65025.2 [51]
La0.95Ag0.05MnO3/
foil Fe2Cr20Al5
27.41 vol.% CH4 in air,
64,410 cm3/(h gcat), 25.7 g catalyst
n.s.n.s.[80]
La0.9Ag0.1MnO3/
foil Fe2Cr20Al5
30.9n.s.n.s.
La0.8Ag0.2MnO3/
foil Fe2Cr20Al5
29.4520n.s.
La0.7Ag0.3MnO3/
foil Fe2Cr20Al5
31.5528n.s.
La0.8Sr0.2FeO34.72 vol.% CH4 in air, 45,000–5000/h54217.8[45]
(La-Ce)FeO35.30.4 vol.% CH4, 2 vol.% O2 in He, 60,000/h, 1.5g catalyst~70026.0[51]
La0.9Ca0.1FeO36.00.4 vol.% CH4, 10 vol.% O2 in N2, 40,000 Ncm3/(h gcat), 0.4 g catalyst54322.6[78]
La0.9Ca0.1FeO314.0n.s.517n.s.[76]
La0.9Ca0.1FeO317.8 505n.s.
La0.9Ca0.1FeO338.037,000 ppmv CH4, 23.22 vol.% O2 in He50897.9[79]
La0.8Ca0.2FeO35.00.4 vol.% CH4, 10 vol.% O2 in N2, 40,000 Ncm3/(h gcat), 0.4 g catalyst53722.6[78]
La0.8Ca0.2FeO38.3n.s.530n.s.[76]
La0.8Ca0.2FeO338.6 503n.s.
La0.8Ca0.2FeO338.037,000 ppmv CH4, 23.22 vol.% O2 in He50295.5[79]
La0.7Ca0.3FeO33.00.4 vol.% CH4, 10 vol.% O2 in N2, 40,000 Ncm3/(h gcat), 0.4 g catalyst52522.6[78]
La0.7Ca0.3FeO35.4n.s.508n.s.[76]
La0.7Ca0.3FeO338.6 505n.s.
La0.7Ca0.3FeO338.037,000 ppmv CH4, 23.22 vol.% O2 in He49494.9[79]
La0.6Ca0.4FeO35.00.4 vol.% CH4, 10 vol.% O2 in N2, 40,000 Ncm3/(h gcat), 0.4 g catalyst54122.6[78]
La0.6Ca0.4FeO38.3n.s.511n.s.[76]
La0.6Ca0.4FeO331.2 511n.s.
La0.6Ca0.4FeO333.037,000 ppmv CH4, 23.22 vol.% O2 in He48794.9[79]
La0.5Ca0.5FeO30.70.4 vol.% CH4, 10 vol.% O2 in N2, 40,000 Ncm3/(h gcat), 0.4 g catalyst63621.4[78]
(Dy-Y)FeO38.60.4 vol.% CH4, 2 vol.% O2 in He, 60,000/h, 1.5 g catalyst~75031.3[51]
(Dy-Y)NiO39.5~67028.4
a T100—the temperature corresponding to 100% methane conversion; b n.s.—not specified.
Table 11. Performances of B-site doped perovskite-type materials for the catalytic combustion of methane.
Table 11. Performances of B-site doped perovskite-type materials for the catalytic combustion of methane.
CatalystSSA (m2/g)Reaction Conditions T50Ea (kJ/mol)Ref.
LaCr0.9Mg0.1O3~5–71.5 vol.%, 18 vol.% in He, 1.2 cm3/s641n.s. a[47]
LaCr0.8Mg0.2O3 647n.s.
LaCr0.7Mg0.3O3 594n.s.
LaCr0.6Mg0.4O3 562n.s.
LaCr0.5Mg0.5O3 553n.s.
LaCr0.5Mg0.5O36.082 vol.% CH4, 18 vol.% O2 in He, 50 Ncm3/min, 0.5 g catalyst577n.s.[48]
LaCr0.5Mg0.5O3∙2MgO13.6552n.s.
LaCr0.5Mg0.5O3∙6MgO24.2545n.s.
LaCr0.5Mg0.5O3∙17MgO36.9529n.s.
LaAl0.95Mn0.05O38.00.4 vol.% CH4, 10 vol.% O2 in N2, 40,000 cm3/(h gcat), 0.4 g catalyst~60728.2[55]
LaAl0.9Mn0.1O37.0n.s.n.s.
LaAl0.8Mn0.2O325.0~54226.1
LaAl0.6Mn0.4O325.0~50725.0
LaAl0.4Mn0.6O326.0~50024.4
LaAl0.2Mn0.8O333.0~44222.8
LaMn0.8Cu0.2O319.00.4 vol.% CH4, 10 vol.% O2 in N2, 40,000 Ncm3/(h gcat)~832n.s.[50]
LaMn0.6Cu0.4O314.0~827n.s.
La(MnPd)O3 (2.32 wt.% Pd)321 vol.% CH4, 4 vol.% O2 in He, 60,000/h542n.s.[67]
La(MnPd)O3 (2.11 wt.% Pd)39482n.s.
LaFe0.9Mg0.1O34.30.4 vol.% CH4, 10 vol.% O2 in N2, 40,000 Ncm3/(h gcat)54723.39[77]
LaFe0.8Mg0.2O35.554123.39
LaFe0.7Mg0.3O37.955223.39
LaFe0.6Mg0.4O39.756525.55
LaFe0.5Mg0.5O35.357925.55
LaFe0.84Cu0.16O34.03.2 vol.% CH4, 12.8 vol.% O2 in Ar, 73.5 mL/min51287[49]
La(Fe,Pd)O3 (2.28 wt.% Pd)221 vol.% CH4, 4 vol.% O2 in He, 60,000/h545n.s.[67]
La(Fe,Pd)O3 (1.25 wt.% Pd)27584n.s.
La(Fe,Pd)O3 (2.47 wt.% Pd)1.6584n.s.
La(Fe,Pd)O3 (2.4 wt.% Pd)14565n.s.
LaCo0.8Cu0.2O321.00.4 vol.% CH4, 10 vol.% O2 in N2, 40,000 Ncm3/(h gcat)~607n.s.[50]
LaCo0.6Cu0.4O313.0~630n.s.
SrTi0.8Zr0.1Mn0.1O3151 vol.% CH4 in air,
~50,000 cm3/(h gcat)
~587n.s.[75]
a n.s.—not specified.
Table 12. Chemical structures of some organic pollutants.
Table 12. Chemical structures of some organic pollutants.
No.CompoundAbbreviationRef.
1Rhodamine B C28H31ClN2O3 479.01 g mol−1
Materials 13 05555 i001
Rh B[290]
2Methyl orange C14H14N3NaO3S 327.33 g mol−1
Materials 13 05555 i002
MO[291]
3Methylene blue C16H18ClN3S 319.85 g mol−1
Materials 13 05555 i003
MB[282]
4Congo red C32H22N6Na2O6S2 696.665 g mol−1
Materials 13 05555 i004
CR[292]
5Neutral red C15H17N4 288.78 g mol−1
Materials 13 05555 i005
NR[293]
6Phenol red C19H14O5S 354.38 g mol−1
Materials 13 05555 i006
Ph R[287]
74-Methyl phenol C7H8O 108.14 g mol−1
Materials 13 05555 i007
p-Cresol[294]
8Tetracycline C22H24N2O8 444.435 g mol−1
Materials 13 05555 i008
TC[295]
Table 13. The performances of complex oxides having perovskite-type structure in the photodegradation of organic dyes.
Table 13. The performances of complex oxides having perovskite-type structure in the photodegradation of organic dyes.
Photocatalyst/
Photoelectrode
DyeLight Source/Intensity Experimental ConditionsDegradation Efficiency (%)Ref.
SrTiO3Rhodamine B (~5 ppm)Ultraviolet (UV) light (200–400 nm)/3 × 15 W tubes 100 mg photocatalyst; Irradiation time: 1.3 h100[301]
SrTiO3Rhodamine B (5 ppm)UV light/15 W226 mg/L photocatalyst; 20 °C; Irradiation: 4.5 h60[302]
SrTiO3Rhodamine B (~5 ppm)Visible (Vis) light (λ > 420 nm)/300 W100 mg photocatalyst in 100 mL; Irradiation time: 6 h<50[304]
Fe-doped SrTiO3 ~85
Nb-doped SrTiO3Rhodamine B (10 ppm)Vis light (λ > 420 nm)Irradiation time: 3 h~50[323]
KNbO3Rhodamine B (40 ppm)UV light/300 W30 mg photocatalyst in 200 mL; Irradiation time: 4 h71[324]
NaNbO3Rhodamine B
(2.5 ppm)
UV light/300 WIrradiation time: 1 h72[325]
LaCoO3Rhodamine B
(2 ppm)
UV light/500 W10 mg photocatalyst; 35 °C; Irradiation time: 0.8 h~100[305]
GdFeO3Rhodamine B (10 ppm)Vis light (λ > 400 nm)/150 W100 mg photocatalyst in 100 mL; Irradiation time: 3 h~90[326]
SmFeO3~95
BiFeO3Rhodamine B (10 ppm)Vis light/100W300 mg photocatalyst; Irradiation time: 3 h >30[327]
BiFeO3Rhodamine B (~5 ppm)Vis light (λ > 420 nm)/500 W100 mg photocatalyst in 50 mL; Irradiation time: 6 h78[328]
BiFeO3Rhodamine B (10 ppm)Vis light (λ > 420 nm)/300 W50 mg photocatalyst in 50 mL; Irradiation time: 6 h~60[329]
Gd-doped BiFeO3Rhodamine B (5 ppm)Vis light (λ > 420 nm)/500 W40 mg photocatalyst in 40 mL; Irradiation time: 2 h94[309]
LaFeO3Rhodamine B (1000 ppm)Vis light (λ > 400 nm)/150 W100 mg photocatalyst in 100 mL; Irradiation time: 3 h100[316]
LaFeO3Rhodamine B (10 ppm)Vis light (λ > 400 nm)/150 W100 mg photocatalyst in 100 mL; Irradiation time: 3 h~96[326]
LaFeO3Rhodamine B (~5 ppm)Vis light (λ > 400 nm)/500 W10 mg photocatalyst; RT; Irradiation time: 2 h76[315]
LaFeO3Rhodamine B (1000 ppm)Vis light (λ > 400 nm)/150 W100 mg in 100 mL; Irradiation time: 12 h93[316]
Ag/LaFeO3Rhodamine B (10 ppm)UV Vis light/125 W100 mg photocatalyst; RT; Irradiation time: 2 h92.8[330]
SrTiO3Methyl orange
(10 ppm)
UV light/15 W75 mg photocatalyst; RT; Irradiation time: 3 h100[303]
Nb-doped SrTiO3Methyl orange
(10 ppm)
Vis light (λ > 420 nm)Irradiation time: 3h~40[323]
LaCoO3Methyl orange
(100 ppm)
Vis light100 mg photocatalyst in 100 mL; Irradiation time: 2h~60[306]
LaCoO3Methyl orangeUV light/30 WIrradiation time: 1.6h89 [331]
BiFeO3Methyl orange
(15 ppm)
UV-Vis light/300 W30 mmol/L photocatalyst; Irradiation time: 8 h>90[307]
BiFeO3Methyl orange
(15 ppm)
Vis light/300 W30 mmol/L photocatalyst; Irradiation time: 16 h>90[307]
BiFeO3Methyl orange
(5 ppm)
Vis light (λ > 420 nm)/300 W200 mg photocatalyst; RT; Irradiation time: 4 h>30[38]
LaFeO3Methyl orange
(10 ppm)
Vis light(λ > 420 nm)/500 WRT; Irradiation time: 4 h>90[317]
Nb-doped SrTiO3Methylene blue
(10 ppm)
Vis light (λ > 420 nm)Irradiation time: 1.3 h~85[323]
KNbO3Methylene blue
(~13 ppm)
Vis light (λ > 420 nm)/180 mW/cm2Irradiation time: 2 h~50[332]
NaNbO3Methylene blueUV light (306 nm)/1mW/cm2Irradiation time: 24 h~15[320]
LaCoO3Methylene blue
(10 ppm)
UV light/30 W200 mg/L photocatalyst; Irradiation time: 1.6 h87[333]
SrFeO3Methylene blue
(~4 ppm)
Vis light/8 WIrradiation time: 12 h100[321]
LaFeO3Methylene blue
(10 ppm)
Vis light(λ>420 nm)/500 WRT; Irradiation time: 4 h93.8[317]
Li-doped LaFeO3Methylene blue
(~31 ppm)
UV Vis light/250 W100 mg photocatalyst in 50 mL; Irradiation time: 1 h45.7[334]
BiFeO3Congo red (20 ppm)Vis light (λ > 420 nm)/500 W2 g/L photocatalyst; RT; Irradiation time: 3 h~40[335]
BiFeO3Congo red (10 ppm)Vis light (λ > 400 nm)/500 WRT; Irradiation time: 4 h~15[336]
Ba-doped BiFeO3Congo red (100 ppm)Vis light/500 WIrradiation time: 120 min~30[337]
Mn-doped BiFeO3Congo redVis light(λ > 400 nm)/ 500 WRT; Irradiation time: 2 h~40[338]
LaCoO3Neutral redUV light/30 WIrradiation time: 0.6 h88[331]
La-doped BiFeO3Phenol red (3.5 ppm)Vis light (λ > 400 nm)/300 W100 mg photocatalyst; RT; Irradiation time: 2 h90.1[333]
LaFeO34-methyphenol
(10 ppm)
Vis light/1500 W1200 mg photocatalyst;
Irradiation time: 6 h
>90[314]
Ca-doped LaFeO3 ~70
Fe-doped SrTiO3Tetracycline
(10 ppm)
Vis light (λ > 420 nm)/300 W100 mg photocatalyst in 100 mL; Irradiation time: 1.3h~71.6[289]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Andrei, F.; Zăvoianu, R.; Marcu, I.-C. Complex Catalytic Materials Based on the Perovskite-Type Structure for Energy and Environmental Applications. Materials 2020, 13, 5555. https://doi.org/10.3390/ma13235555

AMA Style

Andrei F, Zăvoianu R, Marcu I-C. Complex Catalytic Materials Based on the Perovskite-Type Structure for Energy and Environmental Applications. Materials. 2020; 13(23):5555. https://doi.org/10.3390/ma13235555

Chicago/Turabian Style

Andrei, Florin, Rodica Zăvoianu, and Ioan-Cezar Marcu. 2020. "Complex Catalytic Materials Based on the Perovskite-Type Structure for Energy and Environmental Applications" Materials 13, no. 23: 5555. https://doi.org/10.3390/ma13235555

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

Article Metrics

Back to TopTop