Next Article in Journal
Evaluation of Catalysts for the Metathesis of Ethene and 2-Butene to Propene
Previous Article in Journal
Nano-Biochar as a Sustainable Catalyst for Anaerobic Digestion: A Synergetic Closed-Loop Approach
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Heterogeneous Activation of Persulfate by LaMO3 (M=Co, Fe, Cu, Mn, Ni) Perovskite Catalysts for the Degradation of Organic Compounds

by
Donatos Manos
1,
Foteini Papadopoulou
1,
Antigoni Margellou
1,
Dimitrios Petrakis
1,* and
Ioannis Konstantinou
1,2,*
1
Department of Chemistry, University of Ioannina, 45100 Ioannina, Greece
2
Institute of Environment and Sustainable Development, University Research Center of Ioannina (URCI), 45110 Ioannina, Greece
*
Authors to whom correspondence should be addressed.
Catalysts 2022, 12(2), 187; https://doi.org/10.3390/catal12020187
Submission received: 17 December 2021 / Revised: 27 January 2022 / Accepted: 29 January 2022 / Published: 2 February 2022
(This article belongs to the Section Environmental Catalysis)

Abstract

:
Sulfate radical-based advanced oxidation processes (SR-AOPs) are lately applied for the degradation of various pollutants through the formation of reactive oxidant species (ROS) from activation of oxidants, such as persulfate (PS) or peroxymonosulfate (PMS). In this study, LaMO3 (M=Co, Fe, Cu, Mn, Ni) perovskite catalysts were synthesized, characterized by several techniques, and tested for the activation of persulfate towards the degradation of phenolic pollutants. The effect of substitution of position B of La-based perovskites as well as calcination temperature was studied. Overall, the results showed that the decisive role in the catalytic activity was the presence of structures that enhance the transfer of electrons between perovskite and oxidant. LaNiO3 followed by LaCoO3 were found as the most active catalysts. Finally, the stability of the catalysts was studied, showing that B-metal leaching is significant for both catalysts, with LaCoO3 being the most stable one.

1. Introduction

Since water resources’ pollution is one of the major environmental problems in recent years, environmental and chemical technology is moving towards the development of novel and effective decontamination technologies [1]. This need lies in the fact that most emerging organic pollutants, such as pharmaceuticals, antibiotics, pesticides, etc., are resistant, toxic, and exhibit low biodegradability. As a result, it is practically difficult degrade them with conventional methods [2,3,4]. As effective methods of decontamination of water and wastewater, advanced oxidation processes (AOPs) are widely applied for the degradation of various resistant pollutants through the formation of reactive oxidant species (ROS), such as hydroxyl (OH) or sulfate radicals (SO4) [5].
A wide variety of AOPs, such as electrochemical [6], photochemical [7], sonochemical [8], photocatalytic, and catalytic processes [9], can be used for the formation of ROS. These processes may differ both in the way in which ROS are formed as well as in the ROS formed. Nevertheless, they can achieve complete or partial degradation of the pollutants in degradation products of much lower toxicity compared to the original ones [10]. Lately, sulfate radical-based advanced oxidation processes (SR-AOPs) are applied for the degradation of various resistant pollutants through the activation of oxidants, such as persulfate (PS) or peroxymonosulfate (PMS), for the formation of ROS [11]. The formation of sulfate radicals is of particular interest to the scientific community, mainly because of the advantages they exhibit over hydroxyl radicals. Some of the main disadvantages of the decontamination methods based on OH radicals are the limited pH range (2–4) and the creation of sludge in the case of photo-Fenton process [12] and generally the shorter half-life of OH radicals [13]. The characteristics of sulfate radicals that make them more attractive in decontamination methods are mainly the higher redox potential (2.5–3.1 V for SO4 vs. 2.8 V for OH), broader pH range (2–8) of activity [14,15], longer half-life (30–40 μς) [16], and the selectivity towards characteristic groups bearing the pollutant toxicity [17].
Τhe main processes that lead to the generation of sulfate radicals or other ROS are through PS and PMS activation. On the one hand, PMS appears as a triple-potassium salt with the formula 2KHSO5·KHSO4·K2SO4, the so-called Oxone, which is a crystalline solid [18]. Of the three sulfur-containing salts in this formula, the two of them appear as not active for the activation of PMS towards the formation of sulfate radicals [19]. However, PMS is non-toxic, highly soluble in water, and, compared to other oxidants, particularly inexpensive [19]. On the other hand, PS, unlike PMS, displays a symmetrical structure, and the most common forms that can be encountered are potassium-PS (K2S2O8), sodium-PS (Na2S2O8), or ammonium-PS ((NH4)2S2O8), all colorless or white crystalline solids. In the case of decontamination technologies, potassium or sodium salts are mostly used because of their high solubility, while the use of ammonium salt is avoided due to the secondary contamination resulting from the formation of ammonia [20]. Activation of these oxidizing agents can be carried out using several methods, with each method showing a different mechanism with its advantages and disadvantages, e.g., by electrochemical activation [21], UV-irradiation [22], heat [23,24], transition metal catalysts [25], and metal-free catalysts, such as carbon catalysts [26]. Among these methods, the one that appears most effective is heterogeneous catalysis, mainly due to the low energy requirements, the facile separation after the treatment, and the reusability for several catalytic cycles [27,28]. In contrast, in the case of homogeneous catalysis, a major problem that occurs is the difficult separation of the catalyst, resulting frequently in secondary contamination [29].
Catalysts that have been studied in recent years for PS activation are mainly transition metals and transition metal oxides [30,31]. Particular interest in recent years in the activation of PS seems to be shown by perovskite metal oxides. This is a class of materials with the general formula ABO3, where A and B can be occupied by metal cations and, specifically, A being the larger-sized rare earth or alkaline earth metal, B a 3d transition metal, and O is oxygen [32]. Based on the general structure of perovskite, oxygen is the connecting link between the A cation, which is situated at the corners of this cubic structure, and the smaller B cation, which is situated at the center [33]. However, the cubic structure is not the only one that occurs in perovskites, and depending on the cations that occupy the A and B positions, other structures, such as rhombohedral, hexagonal, tetragonal, and orthorhombic ones, can be formed [34]. Perovskites, in addition to the simple ABO3 form mentioned before, can exist in other forms as well, such as A2BO4 (layered perovskite), A2BB′O6 (double perovskite), or A2A′B2B′O9 (triple perovskite), which show even more complex structures [34]. These are materials that have the potential to incorporate many different metal cations into their structure and substitute the cations of positions A and B without destroying their original structure. Still, due to this structure, they are able to incorporate more than 90% of the metallic elements found in the periodic table, while by differentiating the synthesis method, it is possible to obtain materials with different physicochemical properties [35]. Consequently, all these features make perovskites attractive in a multitude of applications, such as sensors and biosensors [36], solar cells and solid oxide fuel cells [37], adsorbents [38], catalysis [39], etc.
Many studies have been presented on the use of perovskites in decontamination technologies through oxidation processes, with emphasis being placed not only on the presence of the various metal ions at the A and B sites of the perovskite structure but also on how these materials are synthesized. For example, the prevalence of Co2+ as the metal ion of position B has been reported in the degradation of organic pollutants through oxidation processes, while studies have also been presented on Fe2+ [39,40], Mn2+ [41,42], Ni2+ [43], Cu2+ [44], and others. However, regarding the position A of the perovskite structure, it seems that lanthanum has prevailed, as the plethora of research studies that have appeared in recent years mainly concern lanthanum-based perovskites.
In this paper, we focus on the synthesis by the surfactant combustion method of a series of lanthanum-based perovskites with different metal ions in position B of the perovskite structure (LaMO3, M=Co, Fe, Cu, Mn, Ni). The synthesized perovskites were characterized by several techniques, and their catalytic activity in terms of the degradation of phenolic compounds by activation of PS was subsequently studied. In the catalytic activity experiments, phenol was used as a standard pollutant, as it belongs to a class of compounds that not only shows high levels of toxicity and has harmful effects on the human body but also on that of animals [45]. In addition, for the most active material, the effect of the calcination temperature during the synthetic method on the catalytic activity was also studied.

2. Results and Discussion

2.1. Characterization of LaMO3

The x-ray diffraction patterns of synthesized perovskite materials are presented in Figure 1. First, diffractograms of LaMnO3 and LaFeO3 materials show the formation of well crystalline perovskite structures with LaFeO3 presenting orthorhombic phase while LaMnO3 presents triangular and cubic perovskite phase. These findings are in agreement with analogous materials synthesized by Rao et al. 2018 [46] and Zhang et al. 2018 [47] for LaFeO3 and Wang et al. 2019b [48] for LaMnO3 although different synthesis approaches have been used, i.e., sol-gel citric acid method for LaFeO3 and hydrothermal-PVP method for LaMnO3.
For LaCoO3, LaNiO3, and LaCuO3, the diffractograms showed a diversity in the crystalline phases formed. LaCoO3 presented three phases, with perovskite monoclinic structure being the more abundant one, followed by orthorhombic perovskite Ruddlesden–Popper (RP) structure and cobalt oxide with cubic structure. In the cases of Hammouda et al. 2017 [49] and Guo et al. 2020 [50], where citric acid sol-gel method was used, cubic and rhombohedral structures occurred, respectively. LaNiO3 also presents three phases, with a greater percentage of orthorhombic RP and smaller percentages of homologous perovskite structure with orthorhombic structure and oxide with triangular structure. In the case of synthesized perovskite with citric acid sol-gel method [40], a cubic structure is reported. LaCuO3 consists of two phases, the main one being orthorhombic RP and, in a smaller percentage, monoclinic oxide, while in the case of Lin et al. (2017) [43], where glycine-combustion preparation method was followed, peaks corresponding to oxides are presented to a small extent, and the main structure corresponds to the rhombohedral structure. Based on the XRD diagram, LaCuO3 perovskite exhibits a higher crystallinity degree as compared to the other catalysts. Finally, for the LaNiO3 material that was also calcined at 800 °C (LaNiO3-800), it was found that the Ruddlesden–Popper perovskite structure decreased, and the largest percentage is found in the classical ABO3 structure. Table 1 shows details of the crystalline phases, as identified by the Match program. The RP designation next to the phases corresponds to RP-perovskite from the Ruddlesden–Popper homologous perovskite series.
Nitrogen adsorption-desorption isotherms at 77 K for all studied materials are shown in Figure 2. For LaMnO3, LaFeO3, LaCoO3, and LaNiO3, the isotherms obtained are of type IV with a hysteresis loop of type H2(b) according to IUPAC classification [51] corresponding to mesoporous materials. The specific surface area ranged from 7 m2/g for LaCoO3 to 14 m2/g for LaFeO3, which is a similar range to several perovskites in the literature [50,52] although there are also cases where synthesized perovskites display specific surface areas in the order of 60–120 m2/g [53,54], while the respective pore diameters were 18.4 Å and 88.9 Å. On the other hand, LaCuO3 and LaNiO3-800 materials appear as non-porous and adsorption-desorption isotherms are of type II with H2(b) type hysteresis loop. Their specific surface areas display values of 7 m2/g and 5 m2/g, respectively. Hydrodynamic particle sizes have been also measured by dynamic light scattering (DLS). Median particle diameters were 0.221, 0.250, 0.288, 0.392, and 2.018 μm for LaNiO3, LaMnO3, LaFeO3, LaCoO3, and LaCuO3, respectively.
Scanning electron microscopy (SEM) was used to reveal the morphology of synthesized perovskites. Based on the microphotographs from all samples for various magnifications (Figure 3), it was observed that none of the materials show a clear geometric crystalline shape. LaMnO3 appears to have a uniform spongy texture, and LaFeO3 appears to have a partly spongy texture, while its surface appears to be rougher than LaMnO3. Moreover, LaCoO3, is also partly spongy, but the interstices between them appear to be more spaced than in the previous cases. Τhe above texture characteristics are consistent with results reported elsewhere [40]. LaNiO3 appears to exhibit a spongy texture as well, which was retained after calcination to 800 °C. LaCuO3 seems to be more “fragmented,” while at the same time, there are larger aggregates in size, and the expected interstices that would make it spongy are not observed.

2.2. Activation of Persulfate by LaMO3 towards Phenolics Degradation

2.2.1. Effect of B-position Metal Ion and Calcination Temperature

According to Figure 4, depending on the metal cation that occupies the B position of each perovskite catalyst, a significant differentiation was observed in the degradation rate of phenolics. The catalytic activity followed the trend: LaNiO3 > LaCoO3 > LaCuO3 > LaMnO3 > LaFeO3, with the corresponding rate constants of 0.082 min−1, 0.029 min−1, 0.006 min−1, 0.0009 min−1, and 0.00004 min−1 (Table 2). Apparently, there is no correlation between the specific surface area of perovskites and their catalytic activity. On the other hand, the presence of two different oxidation states of the metal in position B in the structure of perovskites appears to play a decisive role in their catalytic activity, making easier electron transfer processes between the surface of the perovskites and the oxidant, as it has been reported also elsewhere [42]. For example, in the case of LaCuO3 (La2CuO4-RP-orthorombic 80%), it has been reported that copper is present as Cu(I) and Cu(II) [55]. In addition, RP-perovskite structures showed excellent variability in lattice oxygen and charge-transfer ability. The co-presence of RP and single-perovskite structure was demonstrated to exhibit superior activity for phenol degradation as compared to pure-phase perovskite materials. The higher activity was attributed to the higher amount of generated oxygen vacancies, the stronger affinity of perovskite surface to the reactants (PMS and phenol), and the higher electron conduction efficiency [56]. In consistency, the catalysts presenting mixed RP and single-perovskite structures were the more efficient ones in the present study.
Since LaNiO3 was the most active material in terms of phenolic degradation, the effect of calcination temperature was studied at a higher calcination temperature 800 °C (vs. 600 °C). As can be seen in Figure 4, the catalytic activity of LaNiO3-800 is reduced compared to that calcined at 600 °C. The increase in calcination temperature led to a material that retained its original spongy structure but with a lower specific surface area, the lowest of all previous materials, and a significant variation of the % phase content, with single-LaNiO3 phase prevailing. As a result, the ratio of RP to single-perovskite structure plays a significant role in the catalytic efficiency. However, LaNiO3-800 is still more active than all the other materials with a different B-cation, which reinforces the consideration that crystal phase ratio and the co-current presence of phases with different oxidation states of B-metal is the crucial parameter affecting the activity.
Table 3 summarizes previous data for phenol degradation through persulfate activation with perovskite catalysts. For comparison to the most active materials presented in this research, it should be taken into consideration the different method of substrate determination. The previous studies [56,57,58,59] monitored the degradation of initial phenol concentration by HPLC, while the current method monitored the removal of the total phenolic content (phenol plus phenolic transformation products formed during the process). LaCoO3 displays comparable efficiency (k = 0.029 min−1) to the lower values reported among the other materials. On the other hand, LaNiO3 displays the second best rate constant (0.082 min−1) (Table 3).

2.2.2. Quenching Experiments for the Determination of ROS

Three different scavengers were used in order to elucidate the degradation mechanism and the dominant ROS generated during the treatment. Tert-butanol (ΤΒA), which does not contain α-hydrogens, was used as a scavenger for the generated OH, as it reacts rapidly with the produced hydroxyl radicals. Ethanol (ΕtOH), which contains α-hydrogens, was used as scavenger for both hydroxyl and sulfate radicals [60], and NaN3 was used as a scavenger for both OH and SO4•− radicals as well as singlet oxygen (1O2) [61]. Since the predominance of the synthetic method with the calcination temperature at 600 °C versus 800 °C was presented, the materials that will be studied in terms of the degradation mechanism of phenolics are the most active of this series, namely LaCoO3 and LaNiO3.
As mentioned above, in the case of the LaCoO3/PS system, phenol degradation of 96% was observed in just 90 min, with a corresponding rate constant of 0.029 min−1. In the presence of both ΤΒA and ΕtOH, a minimal decrease in phenol degradation was observed, as shown by the Figure 5a, to 89% and 87% over 90 min, respectively, which means that the formation of OH contributes in a greater percentage to the degradation of phenolics and, in a smaller percentage, of SO4•−. However, in the case when NaN3 was used as a scavenger, a decrease in phenolic degradation was observed to a greater extent. Particularly, a decrease in phenol degradation was observed from 96% in the absence of a scavenger to 82% in 90 min, with the corresponding rate constant decreasing from 0.029 min−1 to 0.015 min−1. Thus, in the LaCoO3/PS system, 1O2 and OH had a greater contribution to the degradation of phenolics while SO4•− were formed in a minor amount.
Additionally, in the case of the LaNiO3/PS system, as can be seen from the Figure 5b, the inhibition in the presence of azide ions predominates since there is a decrease in the degradation of phenolics from 96% (k = 0.082 min−1) in the absence of a scavenger to 71% (0.030 min−1) within 30 min. A significant decrease is also observed in the cases where ΤΒA and ΕtOH were used as scavengers. Specifically, in the case of ΕtOH, there was a decrease in phenol degradation to 76% in 30 min, with the corresponding rate constant of 0.042 min−1, while in the case of TBA, there was a decrease to 86% in 30 min with the rate constant of 0.053 min−1. This means that all the three ROS, namely 1O2, OH, and SO4•−, were formed and contributed the degradation of phenolics.

2.2.3. Reusability and Stability of the Best Catalysts

The stability-reusability of LaCoO3 and LaNiO3 catalysts for three catalytic cycles was studied (Figure 6). In the LaCoO3/PS system, it is evident that the catalyst performance is not affected after three catalytic cycles. In the first two cycles, the degradation of phenol remained stable at 96%, while in the third cycle, there was a slight decrease at 89%. However, this is not observed in the case of the LaNiO3/PS system, as there is a gradual decrease in the catalytic activity of perovskite, i.e., from 96% in the first cycle to 90% in the second cycle, followed by a greater decrease in the third cycle, down to 58%.
Metal leaching was studied as well in order to correlate with the reusability of the catalysts since gradual metal release leads to their deactivation [62]. As observed in Figure 7, there is a gradual increase with a rather constant rate in the amount of cobalt lost from LaCoO3 perovskite, with the total concentration after the third cycle being approximately 6.6 mgL−1. In the case of LaNiO3, higher concentrations of nickel ion were detected, i.e., 14.39 mgL−1 for the first cycle to 17.24 mgL−1 for the third cycle. The subsequent gradual increase in the metal losses after each catalytic cycle has been reported to affect the surface properties of perovskites and thus explains the corresponding decrease in their catalytic activity [50]. Metal leaching in these cases can be due to two reasons: either the existence of even a small degree of water solubility of the catalyst (due to the oxides of the metals they may contain) [63] or the formation of water-soluble products that can result from the reaction of the catalyst with the solvent-oxidant system [64]. Thus, comparing these two materials in terms of their stability, the prevalence of LaCoO3 over LaNiO3 is significant.

3. Materials and Methods

3.1. Materials

For perovskite synthesis, nitrates La(NO3)3 · 6H2O, Mn(NO3)2 · 4H2O, Cu(NO3)2 · 2.5H2O, Ni(NO3)2 · 5H2O, Co(NO3)2 · 6H2O (≥98.0%, Fluka, Steinheim, Germany), and Fe(NO3)3 · 9H2O (≥99.0%, Merck, Darmstadt, Germany) were used as cation precursors. Glycolic acid ethoxylate lauryl ether (≥99.0%, Aldrich, St. Louis, MO, USA) was used in the synthesis in the presence of the anionic surfactant, while the pH adjustment was carried out with 1N HNO3 (Fixanal, Fluka, Seelze, Germany). For the perovskite catalytic activity control experiments, phenol (99% purity), sodium azide (NaN3), and Folin–Ciocalteu’s phenol reagent were obtained from Merck (Germany). Methanol (MeOH, ≥99.8%), ethanol (EtOH), tert-butanol (TBA), sodium carbonate anhydrous (Na2CO3, ≥99.8%), and sodium persulfate (Na2S2O8, ≥98.0%) were purchased from Sigma-Aldrich. All chemicals were used without further purification, and distilled water was used in all experiments.

3.2. Synthetic Method of LaMO3

The synthesis of perovskites was carried out in the presence of the anionic surfactant glycolic acid ethoxylate lauryl ether. Initially, 100 mL of 0.05 M solution of anionic surfactant were stirred for 10 min at 25 °C, 1N HNO3 was added to adjust the pH = 1.00, and then, the solution was stirred for 1h. Then, determined volumes of 0.05 M La(NO3)3 · 6H2O were added, and the mixture was left under stirring for 1h. Afterwards, determined volumes of the corresponding metal nitrate hydrate reagents was added and stirred for 24 h. Finally, the obtain solids were dried at 90 °C and then calcinated at 600 °C for 6 h under atmospheric conditions. Best catalyst (LaNiO3) was also calcined at 800 °C for comparison in the catalytic activity.

3.3. Characterization of LaMO3

X-ray diffraction (XRD) patterns of the synthesized materials were obtained using a D8 Bruker Advance diffractometer with a monochromatic radiation Cu-Ka (λ = 1.5406 Å) in the scanning range 10° ≤ 2θ ≤ 90° and scanning rate of 0.01°/s. The samples were measured in powder form after they were calcined. Identification of crystal phases was carried out via the trial version of Match! 3 Software (Crystal impact, Bonn, Germany).
The surface area of the synthesized materials was measured by adsorption-desorption N2. Adsorption-desorption isotherms were obtained with the Autosorb-1 porosimeter at a liquid nitrogen temperature of 77 K, using a 9-mm measuring cell with the corresponding rod. Degassing of materials at 180 °C for 5 h was required before the measurement. The surface morphology of the particles was observed in a Jeol JSM 5600 system operating at 20 kV. Prior to the analysis, the solids were sputtered with gold under argon atmosphere. A Shimadzu SALD-2300 laser diffraction particle size analyzer (Kyoto, Japan) working with dynamic light scattering (DLS) mode was used for the hydrodynamic particle size measurements.

3.4. Catalytic Activity of LaMO3 Materials

Phenolics degradation experiments were carried out in 200-mL conic flasks containing 100 mL of phenol aquatic solution with an initial concentration of 10 mgL−1 and specific amounts of catalyst (LaMO3) under continuous stirring, at room temperature. In order to start the reaction, oxidant (sodium persulfate) was added into the reactor. Then, at predetermined time intervals, 6 mL of the reaction solution were withdrawn, quenched immediately with methanol to terminate the oxidation reaction, and filtered by 0.45-μm membrane filter. Finally, 5 mL of the collected sample were used for the determination of phenolics concentration by Folin–Ciocalteu method [65] using a UV/Vis spectrophotometer (V-360, Jasco, Tokyo, Japan). The reusability of the catalysts with the best activity was studied through catalytic cycle experiments following the previous procedures. At the end of each catalytic cycle, the catalyst was collected and rinsed with distilled water. The stability of the best catalysts was also studied by determining the concentration of the leached B metals (Co, Ni) at the end of each catalytic cycle, utilizing ICP-MS/Laser Ablation technique. The analysis was performed using a NWR213 laser ablation system (Elemental Scientific, Omaha, NE, USA) attached to an iCAP Q-ICP-MS unit (Thermo Scientific, Waltham, MA, USA).

4. Conclusions

The choice of the metal ion at position B of the perovskite structure appears to significantly affect the properties of the synthesized perovskites, including both their surface properties and the catalytic activity. Of the series of perovskites synthesized, the most active for persulfate activation and subsequent degradation of phenolics was LaNiO3 followed by LaCoO3. The co-presence of RP and single-perovskite structure was demonstrated to exhibit superior activity. Still, it was shown that the increase in calcination temperature resulted in lower catalytic activity. Finally, the use of LaCoO3 and LaNiO3 for consecutive catalytic cycles revealed better recyclability and lower metal leaching for LaCoO3. Further techniques aiming at diminishing metal losses should be considered in future research.

Author Contributions

Conceptualization, I.K. and D.P.; methodology, D.P.; formal analysis, D.M., F.P. and A.M.; investigation, D.M., F.P. and A.M.; resources, I.K. and D.P.; writing—original draft preparation, D.M. and I.K.; writing—review and editing, D.M., D.P. and I.K.; visualization, D.M. and F.P.; supervision, I.K. and D.P; funding acquisition, I.K. and D.P. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Data Availability Statement

Data are contained within the article.

Acknowledgments

The authors would like to thank the unit of ICP-MS/Laser Ablation of the Aristotle University of Thessaloniki for providing access to the facilities. They would also like to thank the XRD unit of the University of Ioannina for providing access to the facilities.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Schwarzenbach, R.P.; Escher, B.I.; Fenner, K.; Hofstetter, T.B.; Johnson, C.A.; von Gunten, U.; Wehrli, B. The Challenge of Micropollutants in Aquatic Systems. Science 2006, 313, 1072–1077. [Google Scholar] [CrossRef]
  2. Arman, N.Z.; Salmiati, S.; Aris, A.; Salim, M.R.; Nazifa, T.H.; Muhamad, M.S.; Marpongahtun, M. A Review on Emerging Pollutants in the Water Environment: Existences, Health Effects and Treatment Processes. Water 2021, 13, 3258. [Google Scholar] [CrossRef]
  3. Babuponnusami, A.; Muthukumar, K. A Review on Fenton and Improvements to the Fenton Process for Wastewater Treatment. J. Environ. Chem. Eng. 2014, 2, 557–572. [Google Scholar] [CrossRef]
  4. Wang, H.; Xi, H.; Xu, L.; Jin, M.; Zhao, W.; Liu, H. Ecotoxicological Effects, Environmental Fate and Risks of Pharmaceutical and Personal Care Products in the Water Environment: A Review. Sci. Total Environ. 2021, 788, 147819. [Google Scholar] [CrossRef] [PubMed]
  5. Zhang, T.; Zhu, H.; Croué, J.-P. Production of Sulfate Radical from Peroxymonosulfate Induced by a Magnetically Separable CuFe2O4 Spinel in Water: Efficiency, Stability, and Mechanism. Environ. Sci. Technol. 2013, 47, 2784–2791. [Google Scholar] [CrossRef] [PubMed]
  6. Sun, Z.; Li, S.; Ding, H.; Zhu, Y.; Wang, X.; Liu, H.; Zhang, Q.; Zhao, C. Electrochemical/Fe3+/Peroxymonosulfate System for the Degradation of Acid Orange 7 Adsorbed on Activated Carbon Fiber Cathode. Chemosphere 2020, 241, 125125. [Google Scholar] [CrossRef] [PubMed]
  7. Kilic, M.Y.; Abdelraheem, W.H.; He, X.; Kestioglu, K.; Dionysiou, D.D. Photochemical Treatment of Tyrosol, a Model Phenolic Compound Present in Olive Mill Wastewater, by Hydroxyl and Sulfate Radical-Based Advanced Oxidation Processes (AOPs). J. Hazard. Mater. 2019, 367, 734–742. [Google Scholar] [CrossRef] [PubMed]
  8. Ferkous, H.; Merouani, S.; Hamdaoui, O.; Pétrier, C. Persulfate-Enhanced Sonochemical Degradation of Naphthol Blue Black in Water: Evidence of Sulfate Radical Formation. Ultrason. Sonochem. 2017, 34, 580–587. [Google Scholar] [CrossRef]
  9. Liu, G.; Zhou, Y.; Teng, J.; Zhang, J.; You, S. Visible-Light-Driven Photocatalytic Activation of Peroxymonosulfate by Cu2(OH)PO4 for Effective Decontamination. Chemosphere 2018, 201, 197–205. [Google Scholar] [CrossRef]
  10. Oh, W.-D.; Dong, Z.; Lim, T.-T. Generation of Sulfate Radical through Heterogeneous Catalysis for Organic Contaminants Removal: Current Development, Challenges and Prospects. Appl. Catal. B Environ. 2016, 194, 169–201. [Google Scholar] [CrossRef]
  11. Duan, X.; Su, C.; Miao, J.; Zhong, Y.; Shao, Z.; Wang, S.; Sun, H. Insights into Perovskite-Catalyzed Peroxymonosulfate Activation: Maneuverable Cobalt Sites for Promoted Evolution of Sulfate Radicals. Appl. Catal. B Environ. 2018, 220, 626–634. [Google Scholar] [CrossRef]
  12. Yang, Q.; Choi, H.; Al-Abed, S.R.; Dionysiou, D.D. Iron-Cobalt Mixed Oxide Nanocatalysts: Heterogeneous Peroxymonosulfate Activation, Cobalt Leaching, and Ferromagnetic Properties for Environmental Applications. Appl. Catal. B Environ. 2009, 88, 462–469. [Google Scholar] [CrossRef]
  13. Machulek, A.; Oliveira, C.S.; Osugi, M.E.; Ferreira, V.S.; Quina, F.H.; Dantas, R.F.; Oliveira, S.L.; Casagrande, G.A.; Anaissi, F.J.; Silva, V.O.; et al. Application of Different Advanced Oxidation Processes for the Degradation of Organic Pollutants. In Organic Pollutants—Monitoring, Risk and Treatment; IntechOpen, Ltd.: London, UK, 2013. [Google Scholar] [CrossRef] [Green Version]
  14. Zhang, B.-T.; Zhang, Y.; Teng, Y.; Fan, M. Sulfate Radical and Its Application in Decontamination Technologies. Crit. Rev. Environ. Sci. Technol. 2015, 45, 1756–1800. [Google Scholar] [CrossRef]
  15. Tsitonaki, A.; Petri, B.; Crimi, M.; Mosbæk, H.; Siegrist, R.L.; Bjerg, P.L. In Situ Chemical Oxidation of Contaminated Soil and Groundwater Using Persulfate: A Review. Crit. Rev. Environ. Sci. Technol. 2010, 40, 55–91. [Google Scholar] [CrossRef]
  16. Mahdi Ahmed, M.; Barbati, S.; Doumenq, P.; Chiron, S. Sulfate Radical Anion Oxidation of Diclofenac and Sulfamethoxazole for Water Decontamination. Chem. Eng. J. 2012, 197, 440–447. [Google Scholar] [CrossRef]
  17. Hu, P.; Long, M. Cobalt-Catalyzed Sulfate Radical-Based Advanced Oxidation: A Review on Heterogeneous Catalysts and Applications. Appl. Catal. B Environ. 2016, 181, 103–117. [Google Scholar] [CrossRef]
  18. Wacławek, S.; Grübel, K.; Černík, M. Simple Spectrophotometric Determination of Monopersulfate. Spectrochim. Acta Part A Mol. Biomol. Spectrosc. 2015, 149, 928–933. [Google Scholar] [CrossRef]
  19. Wacławek, S.; Lutze, H.V.; Grübel, K.; Padil, V.V.T.; Černík, M.; Dionysiou, D.D. Chemistry of Persulfates in Water and Wastewater Treatment: A Review. Chem. Eng. J. 2017, 330, 44–62. [Google Scholar] [CrossRef]
  20. Behrman, E.J.; Dean, D.H. Sodium Peroxydisulfate Is a Stable and Cheap Substitute for Ammonium Peroxydisulfate (Persulfate) in Polyacrylamide Gel Electrophoresis. J. Chromatogr. B Biomed. Sci. Appl. 1999, 723, 325–326. [Google Scholar] [CrossRef]
  21. Cai, C.; Zhang, H.; Zhong, X.; Hou, L. Electrochemical Enhanced Heterogeneous Activation of Peroxydisulfate by Fe-Co/SBA-15 Catalyst for the Degradation of Orange II in Water. Water Res. 2014, 66, 473–485. [Google Scholar] [CrossRef]
  22. Huang, Y.-H.; Huang, Y.-F.; Huang, C.-I.; Chen, C.-Y. Efficient Decolorization of Azo Dye Reactive Black B Involving Aromatic Fragment Degradation in Buffered Co2+/PMS Oxidative Processes with a Ppb Level Dosage of Co2+-Catalyst. J. Hazard. Mater. 2009, 170, 1110–1118. [Google Scholar] [CrossRef] [PubMed]
  23. Tan, C.; Gao, N.; Deng, Y.; An, N.; Deng, J. Heat-Activated Persulfate Oxidation of Diuron in Water. Chem. Eng. J. 2012, 203, 294–300. [Google Scholar] [CrossRef]
  24. Ji, Y.; Dong, C.; Kong, D.; Lu, J.; Zhou, Q. Heat-Activated Persulfate Oxidation of Atrazine: Implications for Remediation of Groundwater Contaminated by Herbicides. Chem. Eng. J. 2015, 263, 45–54. [Google Scholar] [CrossRef]
  25. Dulova, N.; Kattel, E.; Trapido, M. Degradation of Naproxen by Ferrous Ion-Activated Hydrogen Peroxide, Persulfate and Combined Hydrogen Peroxide/Persulfate Processes: The Effect of Citric Acid Addition. Chem. Eng. J. 2017, 318, 254–263. [Google Scholar] [CrossRef]
  26. Duan, X.; Sun, H.; Wang, S. Metal-Free Carbocatalysis in Advanced Oxidation Reactions. Acc. Chem. Res. 2018, 51, 678–687. [Google Scholar] [CrossRef]
  27. Miao, J.; Sunarso, J.; Su, C.; Zhou, W.; Wang, S.; Shao, Z. SrCo1−xTixO3−δ Perovskites as Excellent Catalysts for Fast Degradation of Water Contaminants in Neutral and Alkaline Solutions. Sci. Rep. 2017, 7, 44215. [Google Scholar] [CrossRef] [Green Version]
  28. Oh, W.-D.; Lua, S.-K.; Dong, Z.; Lim, T.-T. Performance of Magnetic Activated Carbon Composite as Peroxymonosulfate Activator and Regenerable Adsorbent via Sulfate Radical-Mediated Oxidation Processes. J. Hazard. Mater. 2015, 284, 1–9. [Google Scholar] [CrossRef]
  29. Rodriguez, S.; Vasquez, L.; Costa, D.; Romero, A.; Santos, A. Oxidation of Orange G by Persulfate Activated by Fe(II), Fe(III) and Zero Valent Iron (ZVI). Chemosphere 2014, 101, 86–92. [Google Scholar] [CrossRef] [PubMed]
  30. Ghanbari, F.; Moradi, M. Application of Peroxymonosulfate and Its Activation Methods for Degradation of Environmental Organic Pollutants: Review. Chem. Eng. J. 2017, 310, 41–62. [Google Scholar] [CrossRef]
  31. Wang, J.; Wang, S. Activation of Persulfate (PS) and Peroxymonosulfate (PMS) and Application for the Degradation of Emerging Contaminants. Chem. Eng. J. 2018, 334, 1502–1517. [Google Scholar] [CrossRef]
  32. Zhang, C.; Zhao, P.; Liu, S.; Yu, K. Three-Dimensionally Ordered Macroporous Perovskite Materials for Environmental Applications. Chin. J. Catal. 2019, 40, 1324–1338. [Google Scholar] [CrossRef]
  33. Zhou, D.; Zhou, T.; Tian, Y.; Zhu, X.; Tu, Y. Perovskite-Based Solar Cells: Materials, Methods, and Future Perspectives. J. Nanomater. 2018, 2018, 8148072. [Google Scholar] [CrossRef]
  34. Assirey, E.A.R. Perovskite Synthesis, Properties and Their Related Biochemical and Industrial Application. Saudi Pharm. J. 2019, 27, 817–829. [Google Scholar] [CrossRef] [PubMed]
  35. Grimaud, A.; May, K.J.; Carlton, C.E.; Lee, Y.-L.; Risch, M.; Hong, W.T.; Zhou, J.; Shao-Horn, Y. Double Perovskites as a Family of Highly Active Catalysts for Oxygen Evolution in Alkaline Solution. Nat. Commun. 2013, 4, 2439. [Google Scholar] [CrossRef] [PubMed]
  36. Jia, F.-F.; Zhong, H.; Zhang, W.-G.; Li, X.-R.; Wang, G.-Y.; Song, J.; Cheng, Z.-P.; Yin, J.Z.; Guo, L.-P. A Novel Nonenzymatic ECL Glucose Sensor Based on Perovskite LaTiO3-Ag0.1 Nanomaterials. Sens. Actuators B Chem. 2015, 212, 174–182. [Google Scholar] [CrossRef]
  37. Li, H.; Zhao, Y.; Wang, Y.; Li, Y. Sr2Fe2−xMoxO6−δ Perovskite as an Anode in a Solid Oxide Fuel Cell: Effect of the Substitution Ratio. Catal. Today 2016, 259, 417–422. [Google Scholar] [CrossRef] [Green Version]
  38. Tavakkoli, H.; Yazdanbakhsh, M. Fabrication of Two Perovskite-Type Oxide Nanoparticles as the New Adsorbents in Efficient Removal of a Pesticide from Aqueous Solutions: Kinetic, Thermodynamic, and Adsorption Studies. Microporous Mesoporous Mater. 2013, 176, 86–94. [Google Scholar] [CrossRef]
  39. Wang, G.; Cheng, C.; Zhu, J.; Wang, L.; Gao, S.; Xia, X. Enhanced Degradation of Atrazine by Nanoscale LaFe1-xCuxO3-δ Perovskite Activated Peroxymonosulfate: Performance and Mechanism. Sci. Total Environ. 2019, 673, 565–575. [Google Scholar] [CrossRef]
  40. Gao, P.; Tian, X.; Nie, Y.; Yang, C.; Zhou, Z.; Wang, Y. Promoted Peroxymonosulfate Activation into Singlet Oxygen over Perovskite for Ofloxacin Degradation by Controlling the Oxygen Defect Concentration. Chem. Eng. J. 2019, 359, 828–839. [Google Scholar] [CrossRef]
  41. Miao, J.; Li, J.; Dai, J.; Guan, D.; Zhou, C.; Zhou, W.; Duan, X.; Wang, S.; Shao, Z. Postsynthesis Oxygen Nonstoichiometric Regulation: A New Strategy for Performance Enhancement of Perovskites in Advanced Oxidation. Ind. Eng. Chem. Res. 2020, 59, 99–109. [Google Scholar] [CrossRef]
  42. Li, X.; Li, M.; Ma, X.; Miao, J.; Ran, R.; Zhou, W.; Wang, S.; Shao, Z. Nonstoichiometric Perovskite for Enhanced Catalytic Oxidation through Excess A-Site Cation. Chem. Eng. Sci. 2020, 219, 115596. [Google Scholar] [CrossRef]
  43. Lin, K.-Y.; Chen, Y.-C.; Lin, Y.-F. LaMO3 Perovskites (M=Co, Cu, Fe and Ni) as Heterogeneous Catalysts for Activating Peroxymonosulfate in Water. Chem. Eng. Sci. 2017, 160, 96–105. [Google Scholar] [CrossRef]
  44. Chen, T.; Zhu, Z.; Wang, Z.; Zhang, H.; Qiu, Y.; Yin, D.; Zhao, G. 3D Hollow Sphere-like Cu-Incorporated LaAlO3 Perovskites for Peroxymonosulfate Activation: Coaction of Electron Transfer and Oxygen Defect. Chem. Eng. J. 2020, 385, 123935. [Google Scholar] [CrossRef]
  45. Anku, W.W.; Mamo, M.A.; Govender, P.P. Phenolic Compounds in Water: Sources, Reactivity, Toxicity and Treatment Methods. In Phenolic Compounds—Natural Sources, Importance and Applications; IntechOpen, Ltd.: London, UK, 2017. [Google Scholar] [CrossRef] [Green Version]
  46. Rao, Y.F.; Zhang, Y.; Han, F.; Guo, H.; Huang, Y.; Li, R.; Qi, F.; Ma, J. Heterogeneous Activation of Peroxymonosulfate by LaFeO3 for Diclofenac Degradation: DFT-Assisted Mechanistic Study and Degradation Pathways. Chem. Eng. J. 2018, 352, 601–611. [Google Scholar] [CrossRef]
  47. Zhang, Y.; Xia, Y.; Li, Q.; Qi, F.; Xu, B.; Chen, Z. Synchronously Degradation Benzotriazole and Elimination Bromate by Perovskite Oxides Catalytic Ozonation: Performance and Reaction Mechanism. Sep. Purif. Technol. 2018, 197, 261–270. [Google Scholar] [CrossRef]
  48. Wang, Y.; Wang, L.; Liu, R.; Li, X. Casein Templated Synthesis of Porous Perovskite and Its Application in Visible-Light Photocatalytic Degradation of Methylene Blue. Mater. Sci. Semicond. Process. 2019, 103, 104597. [Google Scholar] [CrossRef]
  49. Ben Hammouda, S.; Zhao, F.; Safaei, Z.; Srivastava, V.; Lakshmi Ramasamy, D.; Iftekhar, S.; Kalliola, S.; Sillanpää, M. Degradation and Mineralization of Phenol in Aqueous Medium by Heterogeneous Monopersulfate Activation on Nanostructured Cobalt Based-Perovskite Catalysts ACoO3 (A=La, Ba, Sr and Ce): Characterization, Kinetics and Mechanism Study. Appl. Catal. B Environ. 2017, 215, 60–73. [Google Scholar] [CrossRef]
  50. Guo, H.; Zhou, X.; Zhang, Y.; Yao, Q.; Qian, Y.; Chu, H.; Chen, J. Carbamazepine Degradation by Heterogeneous Activation of Peroxymonosulfate with Lanthanum Cobaltite Perovskite: Performance, Mechanism and Toxicity. J. Environ. Sci. 2020, 91, 10–21. [Google Scholar] [CrossRef]
  51. Thommes, M.; Kaneko, K.; Neimark, A.V.; Olivier, J.P.; Rodriguez-Reinoso, F.; Rouquerol, J.; Sing, K.S.W. Physisorption of Gases, with Special Reference to the Evaluation of Surface Area and Pore Size Distribution (IUPAC Technical Report). Pure Appl. Chem. 2015, 87, 1051–1069. [Google Scholar] [CrossRef] [Green Version]
  52. Vijayaraghavan, T.; Bradha, M.; Babu, P.; Parida, K.M.; Ramadoss, G.; Vadivel, S.; Selvakumar, R.; Ashok, A. Influence of Secondary Oxide Phases in Enhancing the Photocatalytic Properties of Alkaline Earth Elements Doped LaFeO3 Nanocomposites. J. Phys. Chem. Solids 2020, 140, 109377. [Google Scholar] [CrossRef]
  53. Amiri, O.; Salar, K.; Othman, P.; Rasul, T.; Faiq, D.; Saadat, M. Purification of Wastewater by the Piezo-Catalyst Effect of PbTiO3 Nanostructures under Ultrasonic Vibration. J. Hazard. Mater. 2020, 394, 122514. [Google Scholar] [CrossRef] [PubMed]
  54. Afzal, S.; Quan, X.; Zhang, J. High Surface Area Mesoporous Nanocast LaMO3 (M=Mn, Fe) Perovskites for Efficient Catalytic Ozonation and an Insight into Probable Catalytic Mechanism. Appl. Catal. B Environ. 2017, 206, 692–703. [Google Scholar] [CrossRef]
  55. Chen, H.; Xu, Y.; Zhu, K.; Zhang, H. Understanding Oxygen-Deficient La2CuO4-Δperovskite Activated Peroxymonosulfate for Bisphenol A Degradation: The Role of Localized Electron within Oxygen Vacancy. Appl. Catal. B Environ. 2021, 284. [Google Scholar] [CrossRef]
  56. Zhu, M.; Miao, J.; Guan, D.; Zhong, Y.; Ran, R.; Wang, S.; Zhou, W.; Shao, Z. Efficient Wastewater Remediation Enabled by Self-Assembled Perovskite Oxide Heterostructures with Multiple Reaction Pathways. ACS Sustain. Chem. Eng. 2020, 8, 6033–6042. [Google Scholar] [CrossRef]
  57. Miao, J.; Sunarso, J.; Duan, X.; Zhou, W.; Wang, S.; Shao, Z. Nanostructured Co-Mn Containing Perovskites for Degradation of Pollutants: Insight into the Activity and Stability. J. Hazard. Mater. 2018, 349, 177–185. [Google Scholar] [CrossRef]
  58. Miao, J.; Duan, X.; Li, J.; Dai, J.; Liu, B.; Wang, S.; Zhou, W.; Shao, Z. Boosting Performance of Lanthanide Magnetism Perovskite for Advanced Oxidation through Lattice Doping with Catalytically Inert Element. Chem. Eng. J. 2019, 355, 721–730. [Google Scholar] [CrossRef]
  59. Lu, S.; Wang, G.; Chen, S.; Yu, H.; Ye, F.; Quan, X. Heterogeneous Activation of Peroxymonosulfate by LaCo1-xCuxO3 Perovskites for Degradation of Organic Pollutants. J. Hazard. Mater. 2018, 353, 401–409. [Google Scholar] [CrossRef]
  60. Xu, Y.; Ai, J.; Zhang, H. The Mechanism of Degradation of Bisphenol A Using the Magnetically Separable CuFe2O4/Peroxymonosulfate Heterogeneous Oxidation Process. J. Hazard. Mater. 2016, 309, 87–96. [Google Scholar] [CrossRef]
  61. Luo, X.; Bai, L.; Xing, J.; Zhu, X.; Xu, D.; Xie, B.; Gan, Z.; Li, G.; Liang, H. Ordered Mesoporous Cobalt Containing Perovskite as a High-Performance Heterogeneous Catalyst in Activation of Peroxymonosulfate. ACS Appl. Mater. Interfaces 2019, 11, 35720–35728. [Google Scholar] [CrossRef]
  62. Sádaba, I.; López Granados, M.; Riisager, A.; Taarning, E. Deactivation of Solid Catalysts in Liquid Media: The Case of Leaching of Active Sites in Biomass Conversion Reactions. Green Chem. 2015, 17, 4133–4145. [Google Scholar] [CrossRef] [Green Version]
  63. Okuhara, T. Water-Tolerant Solid Acid Catalysts. Chem. Rev. 2002, 102, 3641–3666. [Google Scholar] [CrossRef] [PubMed]
  64. Sádaba, I.; Ojeda, M.; Mariscal, R.; Fierro, J.L.G.; Granados, M.L. Catalytic and Structural Properties of Co-Precipitated Mg-Zr Mixed Oxides for Furfural Valorization via Aqueous Aldol Condensation with Acetone. Appl. Catal. B Environ. 2011, 101, 638–648. [Google Scholar] [CrossRef]
  65. Cicco, N.; Lattanzio, V. The Influence of Initial Carbonate Concentration on the Folin-Ciocalteu Micro-Method for the Determination of Phenolics with Low Concentration in the Presence of Methanol: A Comparative Study of Real-Time Monitored Reactions. Am. J. Anal. Chem. 2011, 2, 840–848. [Google Scholar] [CrossRef] [Green Version]
Figure 1. XRD patterns of LaΜO3 perovskites.
Figure 1. XRD patterns of LaΜO3 perovskites.
Catalysts 12 00187 g001
Figure 2. N2 adsorption-desorption isotherms of LaMO3 (M=Co, Fe, Cu, Mn, Ni) perovskites.
Figure 2. N2 adsorption-desorption isotherms of LaMO3 (M=Co, Fe, Cu, Mn, Ni) perovskites.
Catalysts 12 00187 g002aCatalysts 12 00187 g002b
Figure 3. SEM microphotographs of (a) LaMnO3; (b) LaFeO3; (c) LaCoO3; (d) LaNiO3; (e) LaCuO3; and (f) LaNiO3-800.
Figure 3. SEM microphotographs of (a) LaMnO3; (b) LaFeO3; (c) LaCoO3; (d) LaNiO3; (e) LaCuO3; and (f) LaNiO3-800.
Catalysts 12 00187 g003
Figure 4. Degradation of phenolics in different reaction systems (catalyst = 0.25 g L−1, PS = 10 mM, phenol = 10 mgL−1).
Figure 4. Degradation of phenolics in different reaction systems (catalyst = 0.25 g L−1, PS = 10 mM, phenol = 10 mgL−1).
Catalysts 12 00187 g004
Figure 5. Degradation of phenolics in the presence of scavengers; (a) LaCoO3/PS and (b) LaNiO3/PS system (catalyst = 0.25 g L−1, PS = 10 mM, phenol = 10 mgL−1, scavenger= 0.1 M).
Figure 5. Degradation of phenolics in the presence of scavengers; (a) LaCoO3/PS and (b) LaNiO3/PS system (catalyst = 0.25 g L−1, PS = 10 mM, phenol = 10 mgL−1, scavenger= 0.1 M).
Catalysts 12 00187 g005
Figure 6. Catalytic performance after three consecutive cycles of (a) LaCoO3/PS and (b) LaNiO3/PS system (catalyst = 0.25 g L−1, PS = 10 mM, phenol = 10 mgL−1).
Figure 6. Catalytic performance after three consecutive cycles of (a) LaCoO3/PS and (b) LaNiO3/PS system (catalyst = 0.25 g L−1, PS = 10 mM, phenol = 10 mgL−1).
Catalysts 12 00187 g006
Figure 7. Metal (Co, Ni) leaching in the reaction media during each catalytic cycle from the most active catalysts.
Figure 7. Metal (Co, Ni) leaching in the reaction media during each catalytic cycle from the most active catalysts.
Catalysts 12 00187 g007
Table 1. Textural properties of the synthesized perovskites LaMO3.
Table 1. Textural properties of the synthesized perovskites LaMO3.
PerovskiteCrystalline Phase (% Content)Specific Surface Area (m2/g)
LaMnO3LaMnO3-P-trigonal (74%)13
LaMnO3-P-Cubic (26%)
LaFeO3LaFeO3-P-orthorombic (100%)14
LaCoO3LaCoO3-P-monoclinic (70%)7
La2CoO4-RP-orthorhombic (18%)
Co3O4-O-cubic (12%)
LaNiO3La2NiO4-RP-orthorhombic (50%)13
La3Ni2O7-HS-orthorhombic (28%)
NiO-O-trigonal (22%)
LaCuO3La2CuO4-RP-orthorombic (80%)7
CuO-O-monoclinic (20%)
LaNiO3-800LaNiO3-P-trigonal (69%)5
La2NiO4-RP-tetragonal (25%)
NiO-O-trigonal (6%)
Table 2. Catalytic activity of the LaMO3 (M=Co, Fe, Cu, Mn, Ni) perovskites.
Table 2. Catalytic activity of the LaMO3 (M=Co, Fe, Cu, Mn, Ni) perovskites.
Perovskite(%) Degradationk (min−1)R2
LaMnO310% in 120 min0.00090.9696
LaFeO35% in 120 min0.000040.9625
LaCoO396% in 90 min0.0290.9258
LaNiO396% in 30 min0.0820.9344
LaCuO352% in 120 min0.0060.9925
LaNiO3-80097% in 45 min0.0480.9248
Table 3. Performance of various perovskite systems for phenol degradation.
Table 3. Performance of various perovskite systems for phenol degradation.
Catalyst(%) Degradationk (min−1)Metal Leaching (mgL−1)Reference
LaCo0.5Mn0.5O3+δ100% in 60 min0.052Co: 4.07[57]
La0.4Sr0.6MnO3-δ100% in 90 min0.031Mn: 1.11
Sr: 8.91
[58]
LaCo0.6Cu0.4O399% in 12 min0.302Co: 1.37[59]
La0.4Sr1.05MnO4-δ100% in 60 min0.070Mn: 3.20
Sr: 4.04
[56]
LaCoO396% in 90 min0.029Co: 1.34This work
LaNiO396% in 30 min0.082Ni: 14.39This work
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Manos, D.; Papadopoulou, F.; Margellou, A.; Petrakis, D.; Konstantinou, I. Heterogeneous Activation of Persulfate by LaMO3 (M=Co, Fe, Cu, Mn, Ni) Perovskite Catalysts for the Degradation of Organic Compounds. Catalysts 2022, 12, 187. https://doi.org/10.3390/catal12020187

AMA Style

Manos D, Papadopoulou F, Margellou A, Petrakis D, Konstantinou I. Heterogeneous Activation of Persulfate by LaMO3 (M=Co, Fe, Cu, Mn, Ni) Perovskite Catalysts for the Degradation of Organic Compounds. Catalysts. 2022; 12(2):187. https://doi.org/10.3390/catal12020187

Chicago/Turabian Style

Manos, Donatos, Foteini Papadopoulou, Antigoni Margellou, Dimitrios Petrakis, and Ioannis Konstantinou. 2022. "Heterogeneous Activation of Persulfate by LaMO3 (M=Co, Fe, Cu, Mn, Ni) Perovskite Catalysts for the Degradation of Organic Compounds" Catalysts 12, no. 2: 187. https://doi.org/10.3390/catal12020187

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