Next Article in Journal
Solar and Convective Drying: Modeling, Color, Texture, Total Phenolic Content, and Antioxidant Activity of Peach (Prunus persica (L.) Batsch) Slices
Next Article in Special Issue
The Preparation of Mn-Modified CeO2 Nanosphere Catalysts and Their Catalytic Performance for Soot Combustion
Previous Article in Journal
Feasibility Study on Space Reorientation for Liquid Hydrogen Tanks by Means of Evaporated Exhaust Gas
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Enhanced Activation of Peroxymonosulfate via Sulfate Radicals and Singlet Oxygen by SrCoxMn1−xO3 Perovskites for the Degradation of Rhodamine B

1
College of Chemical Engineering and Materials Science, Tianjin University of Science and Technology, Tianjin 300457, China
2
Institute of Engineering Technology, Sinopec Catalyst Co., Ltd., Beijing 100029, China
*
Authors to whom correspondence should be addressed.
Processes 2023, 11(4), 1279; https://doi.org/10.3390/pr11041279
Submission received: 23 March 2023 / Revised: 11 April 2023 / Accepted: 18 April 2023 / Published: 20 April 2023
(This article belongs to the Special Issue Metal-Support Interactions in Heterogeneous Catalysis)

Abstract

:
Perovskite is of burgeoning interest in catalysis, principally due to such material having high thermal stability, modifiable variability, ferromagnetism, and excellent catalytic performance in peroxomonosulfate (PMS) activation. In this study, the SrCoxMn1−xO3 perovskites with different Mn doping were synthesized by a facile sol-gel method for peroxymonosulfate (PMS) activation to degrade Rhodamine B. The obtained SrCo0.5Mn0.5O3 perovskite exhibited the best catalytic efficiency, as Rhodamine B (40 mg/L) was removed completely within 30 min. In the system of SrCo0.5Mn0.5O3–PMS, several reactive species were produced, among which sulfate radicals and the singlet oxygen mainly contributed to Rhodamine B degradation. The relatively high catalytic performance could be attributed to the coupled redox cycle between Mn and Co, and the abundant oxygen vacancies. Moreover, the SrCo0.5Mn0.5O3 catalyst showed excellent stability and reusability, maintaining a high catalytic activity after several cycling tests. This study demonstrated that the Mn doping of SrCoO3 could not only enhance the B-site activation in SrCo0.5Mn0.5O3 but also enrich the oxygen vacancies, thus improving the efficiency of PMS activation.

1. Introduction

Organically synthetic dyes have been used widely in people’s daily lives with the increase of industrialization, along with the increased demand for products related to fiber-material products, printing, papermaking, medicine, and information materials industries. Dyes are also used to make dye-sensitized solar cells with potentially high photoelectric conversion efficiency to promote a greener energy transition [1,2]. It is noted that these highly stable industries are, by nature, usually toxic and carcinogenic, which poses a great threat to the environment and human health [3,4,5]. As a representative contaminant, Rhodamine B (RhB) (C28H31ClN2O3) has a high water solubility and carcinogenicity hazard [6]. To treat such contaminants, advanced oxidation processes (AOPs) have been investigated widely as an effective approach due to their excellent oxidation performance [7,8,9].
Compared with the traditional AOPs, sulfate radical-based AOPs (SR-AOPs) have gradually become promising technologies in water treatment due to their strong oxidation ability in a wide pH range [10,11,12,13]. Typically speaking, sulfate radicals ( SO 4   · ) generated in the activation of PMS have a higher standard reduction potential (E0 = 2.5–3.1 V) [14,15] and better selectivity for most refractory organic pollutants [16,17]. Though homogeneous catalysts performed excellent PMS activation, the transition metal ions are a serious threat to the ecosystem and a risk of secondary pollution [18]. Hence, the activation of PMS by catalysts was considered a feasible method for the degradation of contaminants due to its economical, reusability, and operational simplicity [19,20]. In addition, it avoided the secondary pollution caused by toxic metal ions to a certain extent. Among these, the perovskite catalysts/PMS system has drawn great attention due to its feasibility to be separated from the water originating from its ferromagnetism, and modifiable variability by its abundant structure modification possibility [21,22,23].
Perovskite has a typical structure of ABO3, and various elements with potent redox properties could be chosen to reside in A or B sites [24], thus providing the feasibility to modify its catalytic property [25]. For simple ABO3 perovskite, it was reported that SrCoO3 and LaCoO3 perovskite exhibited a higher activity under the PMS activation for phenol degradation [26]. Meanwhile, common transition elements such as Co, Mn, and Fe constructed as a B site show high efficiencies towards SR-AOPs [27,28,29]. Co-based perovskites are investigated as a PMS activator with widespread attribution of the redox compatibility and impressive PMS activation by Co species [23,30]. Mn-based perovskites usually show unique performance on PMS activation due to the high surface area, the multivalence nature of Mn ions, and the oxygen defects [31,32].
To further improve the perovskite performance, a third metal cation can be doped in the B site to control the oxidation states of the original element and obtain more oxygen vacancies [33,34,35]. For example, the bimetallic synergistic doping of Co and Fe in LaCo0.5Fe0.5O3 exhibited an impressive removal efficiency of bisphenol A(BPA) [29]. The Cu-substituted LaFeO3 perovskite not only created a metal redox cycle between Fe3+ and Cu+ but also increased the amount of the surface hydroxyl groups [36]. The reactive oxygen species in layered perovskite LaSrCo0.8Fe0.8O4 facilitates the formation of nonfree radical singlet oxygen, which shows exceptionally efficient PMS activation [37]. Therefore, to enhance the PMS activation, based on achieving more oxygen vacancies and accelerating the redox cycling of the transition metals, we tried to synthesize the SrCoxMn1−xO3 perovskites and explore the effect of Mn on the B site.
In this paper, SrCoxMn1−xO3 perovskite was synthesized by the facile sol-gel method. RhB was selected as the target pollutant to examine the activation of PMS under the affection of the catalyst. The reaction conditions such as PMS concentration, catalyst dosage, pH, reusability, and stability were also explored. Furthermore, the mechanism of the interaction between the singlet oxygen and the persulfate radical was confirmed in SrCo0.5Mn0.5O3–PMS system by EPR and quenching experiments. The SrCo0.5Mn0.5O3 perovskite was found to have more oxygen vacancies and show superior catalyst activation and stability with the doping of Mn in this study. The SrCo0.5Mn0.5O3–PMS system would be a promising candidate to degrade RhB in wastewater treatment.

2. Experiment

2.1. Materials

All reagents in the experiment were of analytical grade. Strontium nitrate, tert-butanol, and citric acid monohydrate were from Xi Long Scientific(Beijing, China). Cobalt nitrate hexahydrate (Co(NO3)2·6H2O), and 50% manganese nitrate solution (Mn(NO3)2·50 wt%) were from Tianjin Da Mao Chemical Reagent Factory(Tianjin, China). PMS (2KHSO5·KHSO4·K2SO4), p-benzoquinone, L-histidine, and sodium thiosulfate were from Macklin(Shanghai, China). The Rhodamine B was from Tianjin Fuchen Chemical Reagent Factory(Tianjin, China). The solution pH was adjusted by a 0.1 M H2SO4 or NaOH solution. Deionized water was used throughout the work.

2.2. Preparation of Catalysts

SrCoxMn1−xO3 perovskites were synthesized via the sol-gel method using citric acid as an organic complexing agent. For a typical synthesis of SrCoO3, citric acid, strontium nitrate, and cobalt nitrate hexahydrate were poured into the beaker sequentially and stirred vigorously (450 rpm), then continuously heated to 90 °C until the total evaporation of the solvent. The resulting powder was dried at 110 °C overnight and calcined at 900 °C for 5 h with a heating rate of 2 °C min−1. For the SrCoxMn1−xO3 perovskites, manganese nitrate as another B-site component was added to the mixed solution as a certain molar ratio (x = 0, 0.3, 0.5, 0.6,0.7, 0.8, 0.9, and 1.0, respectively).

2.3. Characterization of Catalysts

X-ray diffraction (XRD) patterns of all perovskite catalysts were characterized by the X-ray diffractometer (Smart Lab, Rigaku, Japan) with Cu Kα line as the source (60 kV, 150 mA) and the scanning range (2θ) was from 10° to 90°. The elemental compositions of the samples were analyzed by X-ray fluorescence spectroscopy (ZSX PrimusIV). Scanning electron microscopy (SEM) was applied to observe the morphology and structure of the perovskites by Gemini SEM 300 (ZEISS, Germany). Energy dispersive spectroscopic (EDS) mapping was also applied to verify the uniformity of the element. To analyze the specific surface areas and the pore-size distributions, nitrogen physisorption was performed on a Micromeritics ASAP 2460 analyzer. The leaching volume of catalysts was determined by ICPE-9820 (Shimadzu, Japan). The concentration of RhB was all obtained by UV-Vis-2600i. The oxygen defect was determined by electron paramagnetic resonance (EPR) at 9.853 GHz by EMXPlus. For O2-TPD (temperature programmed desorption) (AMI-300IR), the perovskites were treated at 500 °C for 1 h in an oxygen atmosphere. Then, the cooled perovskites purged by argon were heated to 900 °C under a constant argon flow, with a flow rate of 30 mL/min and a heating rate of 10 °C min−1. To confirm the synthesis of the perovskites, the FT-IR was measured by Thermo Fisher Nicolet Is5 in a wave number from 400–2000 cm−1. X-ray photoelectron spectroscopy (XPS, Thermo ESCALAB 250XI) was applied to obtain the chemical valence of fresh and used catalysts. All the binding energies were calibrated with C 1 s peak at 284.8 eV. Radicals in the system were identified by the EPR test (Bruker EMXplus) with the DMPO and TEMP as the spin-trapping agents.

2.4. Catalytic Activity Measurements

In the typical experiment, Rhodamine B (RhB) was selected as a target pollutant to evaluate the catalyst activity. The concentration of RhB was tested by a UV-Vis spectrophotometer based on the Lambert-Beer law. Experiments were carried out in a 300 mL beaker with 250 mL 40 mg/L RhB solution with continuous stirring at 300 rpm in the magnetic stirrer. The specific number of SrCoxMn1−xO3 catalysts was added into the breaker and stirred for 30 min to reach adsorption equilibrium. When the concentration of RhB reached equilibrium, certain amounts of PMS were added to the solution. The initial pH (3–9) was adjusted by a 0.1 M H2SO4 or 0.1 M NaOH solution and the temperature was held at 25 °C. At the specified time point, about 2.5 mL of the reaction solutions were withdrawn through a 0.22 μm membrane filter and quenched immediately with 40 μL 0.1 M Na2S2O3 solution. The collected perovskites were analyzed by the UV-VIS-2600i (Shimadzu Japan) to obtain the removal efficiency of the RhB. Different dosages of catalysts and PMS were studied to obtain the optimal reaction conditions. The recycling experiments were also performed to investigate the reusability and stability of the catalyst. The used perovskite catalyst was washed with deionized water and ethanol three times in sequence and dried overnight at 80 °C in the oven. And the leaching of Co and Mn cations was measured by ICP spectrometry. To confirm the dominant reactive radicals in the SrCo0.5Mn0.5O3–PMS system, tert-butanol, p-benzoquinone, methanol, and L-histidine were used as quenching agents for the different active radicals respectively.

3. Results and Discussion

3.1. Characterization of Catalysts

The actual substitution of Mn was analyzed by XRF and listed in Table 1. For the SrCoxMn1−xO3 perovskites, the actual metal content was consistent with the named content. To analyze the crystal structure of the perovskites, an XRD analysis was conducted, as shown in Figure 1a. According to the X-ray powder diffraction patterns of the SrCoxMn1−xO3 perovskites, the SrCoO3 perovskite phase (JCPDS number 49-0692) and the SrMnO3 (JCPDS number 24-1213) were observed [38]. From Figure 1b, the shift of diffraction peaks to a lower angle suggests the dope of Mn inside SrCoOx, which enlarges the unit cell parameters due to its larger ionic radius (Mn3+: 0.66 Å, Co3+: 0.63Å). It indirectly proved that manganese was doped into the perovskite lattice. From Table S1, the lattice parameters decreased with the Mn doping, which may be caused by the formation of anion defects [39]. It is worth highlighting that when the Co:Mn ratio decreased to 3:7, it had an obvious superposition of the diffraction peaks of SrCoO3 and SrMnO3. Meanwhile, the characteristic of SrMnO3 was weaker than SrCoO3, which may be explained by the following two aspects: better affinity for the forming perovskite phase and the Mn phase hinders the decomposition of carbonate [40,41]. These results implied that Mn can be doped into the SrCoO3 perovskite partially and eventually become a mixture with the increase of Mn. Additionally, the detailed SrCoxMn1−xO3 perovskites were characterized by XRD (Figure S1).
As in Table 2, the specific surface area of SrCoO3 was small, which is a typical property for perovskite materials [42,43], which may be attributed to the high synthesis temperature. With the addition of Mn, though the specific surface area of SrCoxMn1−xO3 increased from 0.63 m2/g to 4.21 m2/g, all perovskites demonstrated a small surface area, which presumed the key factors of the difference regarding their catalytic activity was not the surface area. Details of the surface texture of the catalysts are listed in Table S1 and the typical nitrogen adsorption-desorption isotherm is shown in Figure S2, where it can be seen that the SrCoxMn1−xO3 perovskites did not change much in pore volume and pore size.
The particulate morphology of SrCoxMn1−xO3 was examined by SEM. From Figure 2, it can be found that when compared with the Mn-replaced SrCoxMn1−xO3 perovskites, the SrCoO3 perovskites have more severe particle aggregation. EDS element mappings clearly showed a uniform distribution of Sr, Mn, Co, and O on the SrCo0.5Mn0.5O3 (Figure 2g). Moreover, the EDS results obtained from a microzone (Figure S3) are consistent with those obtained by XRF to further prove the uniform distribution at a microscale.

3.2. Active Oxygen Species Analysis

Catalyst-assisted SR-AOPs are related to the activation ability of oxygen-related intermediates in reaction. O2-TPD experiments were applied to study the oxygen species in different Mn-substituted perovskites. As depicted in Figure 3a, the characteristic peaks of oxygen that evolved from the SrCoxMn1−xO3 perovskites mainly appeared at high temperatures (>800 °C), which resulted from the lattice oxygen species [44]. There are no obvious signal peaks of surface weakly adsorbed oxygen from 200 to 500 °C, which corresponded to the very-low specific surface area [45,46]. With the increase of Mn-substitution, the amount of oxygen released from the lattice oxygen gradually increased [47]. As a whole, there was no obvious difference in lattice oxygen among the perovskites.
In addition to the existing oxygen species, metal oxides usually consisted of active oxygen vacancies, which could promote the activation of oxygen-related precursors. Perovskites are typically active in redox reactions partially due to their abundant oxygen vacancies, especially when rationally doped. Oxygen vacancy was characterized by the EPR technique to investigate the variation caused by the dope of Mn. As shown in Figure 3b, the centrally symmetrical signals at the g factor of 2.00 were ascribed to the oxygen vacancies. The strongest signal from SrCo0.5Mn0.5O3 perovskite illustrated that the substitution of Mn led to more oxygen vacancies. The result was attributed to the Mn doping leads distortion in the Co-O octahedra and weaker metal-oxygen bonds, which reduced the formation of oxygen vacancies [40]. In comparison, SrCoO3 and SrMnO3 all showed a lesser amount of oxygen vacancies, which proved the promotion effect of oxygen vacancy generation by Mn doping.

3.3. The Catalytic Activity of Catalysts

As a typical dye contaminant in wastewater, RhB was selected to evaluate the catalytic performance of perovskites for the activation of PMS. From Figure S5, sole PMS could oxidize only 10% of RhB, and the absorption capability of SrCoO3 (without PMS) for RhB was limited (only 10%) in 30 min, meaning the sole PMS or perovskite catalyst could only contribute in a limited way to the removal of RhB. When PMS was added simultaneously with SrCoxMn1−xO3, the concentration of RhB was reduced to close to zero within 25 min (Figure 4a), which showed that all SrCoxMn1−xO3 perovskites had high efficiency for PMS activation.
The catalytic degradation activity of RhB using different SrCoxMn1−xO3 (x from zero to one) activating PMS is shown in Figure 4a. Compared to the pure perovskites with only Mn or Co cations, it can be seen that the B site doped perovskites showed much higher catalytic performance. From Table 3, the degradation kinetics of RhB in different perovskites/PMS was shown by applying the first-order kinetics, which were as below:
ln ( C t / C 0 ) = k t
The catalytic-activity kinetic constant of SrCo0.5Mn0.5O3 (k = 0.306 min−1) was the highest among all the catalysts, with an order of magnitude higher than the undoped SrCoO3 (k = 0.081 min−1) perovskites. Meanwhile, it can be observed that the kinetic constant of RhB degradation on SrCo0.5Mn0.5O3 was 3.78 and 30.6 times that of the SrCoO3 and SrMnO3 perovskites. Therefore, the catalytic performance of SrCo0.5Mn0.5O3 should be attributed to the doping effect of Mn, since the SrMnO3 phase had a poor catalytic performance even worse than its counterpart SrCoO3.
Another factor to consider in the actual environment is the stability of the catalyst. The dissolution of the active centers usually leads to the failure of the catalyst in routine running. The leaching experiment was applied to monitor the stability test. As in Figure 4c, Mn is stable for all the catalysts compared with Co, and the leaching of the cobalt cation in SrCo0.5Mn0.5O3–PMS was only 0.22mg/L after the reaction (30 min), which was less than half of that in the SrCoO3–PMS system (Figure 4c). Therefore, the SrCo0.5Mn0.5O3 catalyst exhibits good catalytic activity and stability.
To further analyze the possible function of the dissolved Co2+, an exact concentration of Co2+ (0.22 mg/L) originated from CoSO4 and Co(NO3)2 was investigated under the same degradation experiment system, without adding any solid catalysts. As shown in Figure 4d, compared to the SrCo0.5Mn0.5O3–PMS system, the simulated leached Co2+ cases show relatively worse reactivity, with only 85% of the RhB removed in 30 min. Taking into account that the actual concentration of Co2+ leached was much less than in the simulated cases, it can be indicated that the heterogeneous reaction had a fundamental effect on the system.
To further investigate the stability and reusability of the catalyst, four consecutive degradation experiments were conducted. As displayed in Figure 5, the removal of RhB maintained 94% at 30 min after the 4th cycle, which exhibited good reusability for SrCo0.5Mn0.5O3 perovskite.
Furthermore, to evaluate the overall performance of SrCo0.5Mn0.5O3–PMS for removing RhB, a comparison of several typical catalysts for PMS activation was summarized in Table 3. In this work, SrCo0.5Mn0.5O3 showed a very high activation ability for PMS and catalyst stability.

3.4. Effect of Reaction Parameters on RhB Degradation

The effect of various operation factors in the SrCo0.5Mn0.5O3–PMS system for the removal of RhB, including PMS concentration, the dosage of SrCo0.5Mn0.5O3, and pH, were investigated to illustrate the relationship between the operation and degradation efficiencies.
As shown in Figure 6a, when the catalyst concentration was increased from 0.015 g/L to 0.030 g/L, it can be observed that the removal efficiency of RhB was significantly enhanced from 69.8% to 100.0% within 30 min. The phenomenon was likely attributed to more active radicals generated from more SrCo0.5Mn0.5O3 perovskite. However, when the catalyst concentration was increased from 0.045 g/L and 0.06 g/L, the degradation rates of RhB maintained almost the same, since PMS might exhibit a self-quenching effect caused by excess metal species and radicals.
The influence of PMS concentration on the degradation efficiency of RhB was further explored. As in Figure 6b, the general trend displayed that the PMS concentration (0.02 mM to 0.04 mM) had a significant enhancement effect on RhB degradation. When the PMS concentration further changed from 0.04 mM to 0.08 mM, the removal efficiency of RhB showed only a slight ascent, from 90% to 100%, within 25 min. It can be ascribed to the self quenching of the generated radicals with PMS [56,57] by Equations (1) and (2).
SO 4   · + HSO 5   HSO 4   + SO 5   ·
· OH + HSO 5   SO 5   · + H 2 O
The initial pH has played a vital role in the degradation of RhB by the SrCo0.5Mn0.5O3–PMS system, as shown in Figure 6c. The perovskites maintained a high catalytic performance as the initial pH ranged from three to seven, though the removal efficiency did not perform well at the initial pH of nine (85.6% within 30 min), which may be attributed to the fact that the generation of SO 4   · was inhibited under alkaline conditions. As shown in Figure S6, the pHzvc of SrCo0.5Mn0.5O3 was 7.83. SrCo0.5Mn0.5O3 perovskite is negatively charged when the intimal pH > 7.83 is a hinderance to the activation of PMS due to the repulsive force [58]. A significant drop in the removal efficiency of RhB in the first three minutes (47%) can be observed at an initial pH of three. This observation can be attributed to the fact that the protonated surfaces on the catalytic are easier to adsorb HSO 5 and generate much more radicals.

3.5. Identification of Mainly Reactive Species in the SrCo0.5Mn0.5O3–PMS System

Overall, the degradation of RhB has two paths: chromophore-structure destruction and deethylation [49]. From the UV-VIS spectra (Figure S7), the characteristic peak did not blue shift or red shift, which showed the cleavage of the whole conjugated chromophore structure of RhB is easier than the deethylation process. Thus, the radicals mainly contributed to the chromophore-structure destruction.
To identify the main reactive radicals, quenching experiments were conducted in the SrCo0.5Mn0.5O3–PMS system. As is commonly suggested in SR-AOPs, four types of reactive radicals could be generated in the activation of PMS, including peroxy-sulfate, hydroxyl radical (·OH), singlet oxygen (1O2), and superoxide radical( O 2   · ) [14]. As shown in Figure 7a, TBA, EtOH, PBQ, and L-histidine were added as the quencher agents of ·OH, SO 4   · along with ·OH, O 2   · , and 1O2, respectively [59,60]. The addition of 0.12 M TBA or 0.06 mM PBQ reduced the RhB removal efficiency slightly from 100% to 97.1% and 96.1%, respectively. It can be surmised that the hydroxyl radicals (·OH) are not the main reactive radicals, and the superoxide radical does not work well. In comparison, the decomposition efficiency of RhB decreased from 100% to 71.4% within 30min when 0.12 M EtOH was added (nEtOH:nPMS = 2000:1). By combining the fact that TBA did not quench the reaction significantly, it can be deduced that it was SO 4   · rather than ·OH contributing to the degradation. Specifically, the addition of 0.06 mM L- histidine exhibited an impressive competitive inhibition effect, which made the RhB degrade only 14.1% within 30 min, suggesting that 1O2 could also be a key intermediate together with SO 4   · . According to the above analysis, the SO 4   · radical and the singlet oxygen had a great effect on the system.
To further verify the results of the quenching effect, the EPR analysis was tested with the DMPO as a spin-trap agent for the generated SO 4   · radical and the ·OH radical, and TEMP for 1O2, respectively [61,62]. As shown in the Figure 7b, the characteristic peaks of the DMPO-·OH adducts (hyperfine splitting constants αH = αN = 14.8 G) and DMPO- SO 4   · H = 0.78 G, αH = 1.48 G, αH = 9.6 G and αN = 13.2 G) in 3 and 12 min were observed in the EPR spectrums, the upward trend in the catalytic process confirmed the sustained generation of active radicals. Meanwhile, the intensity of the DMOH-·OH adduct had a significant enhancement, and the DMOH- SO 4   · had only weak growth, indicating that it was continuously consumed, which caused the slightly left shift with the time [36]. Additionally, as shown in Figure 7c, three-line peaks were detected in the EPR spectrums with the relative intensities of 1:1:1, which matched well with the signals of the TEMP-1O2 adduct.

3.6. Catalytic Mechanism in the SrCo0.5Mn0.5O3–PMS System

To better investigate the reusability of the catalyst, the XPS spectra of the SrCo0.5Mn0.5O3 were analyzed before and after the four-cycles activity test. Characteristic peaks of cobalt, manganese, oxygen, and strontium were obtained from the wide XPS spectra (Figure 8a), in which a minor N 1s signal could be observed in the used catalyst, which could have originated from the decomposition of RhB. The deconvolution results of fresh and used SrCo0.5Mn0.5O3 perovskites were all displayed in Tables S2 and S3. From the high-resolution XPS spectrum of Co 2p (Figure 8b), the two main peaks at 780.10 eV and 795.3 eV represent Co 2p3/2 and Co 2p1/2, respectively, with three shake-up satellites of Co ions at around 789.97 eV and 804.29 eV. The Co 2p1/2 peaks at 780.10 eV and 795.10 eV were recognized as Co3+, with the energy splitting about 15 eV. After the catalytic cycles, the relative intensity of Co2+/Co3+ dropped from 81.01% to 73.73%, which demonstrates that Co2+ played a positive role. The Mn 2p peak had significantly split spin-orbit components at the binding energy of 642.00 eV and 653.2 eV. Through the deconvoluted Mn 2p spectra (Figure 8c), two oxidation states at 642.00 eV (Mn3+) and 644.00 eV (Mn4+) accounted for 68.26% and 31.74% and then changed to 71.91% and 28.09% after the reaction. The increase in the ratio of Mn4+/Mn3+ and Co3+/Co2+ indicated that the redox of the two B-site metals was included in the PMS system.
The high-resolution XPS spectra of O 1 s on SrCo0.5Mn0.5O3 before and after the reaction were displayed in Figure 8d, in which four oxygen species were found after deconvoluting. Four individual peaks at 529.13 eV, 530.20 eV, 531.50 eV, and 532.8 eV could be regarded as lattice oxygen (O2−, OI), oxygen defects ( O 2 2 /O, OII), surface absorbed oxygen species (OIII) and adsorbed molecular H2O(OIV). The deconvolution results showed that the ratio of Olatt had a slight drop of 9.72% and the oxygen species has a sharp decline of 31.09%, suggesting they played an important role in the PMS activation process. Surface-absorbed oxygen species can improve reactive activity by accelerating the electron transport between PMS and activates through the formation of hydrogen bonds [63]. Meanwhile, the absorbed molecular H2O increased by 8.72% more than the initial catalyst, which could be owing to the adsorption of organic intermediates in the process.
From the discussion above, the two PMS activation mechanisms (radical-based and nonradical processes) on the SrCo0.5Mn0.5O3 perovskite could be proposed as follows (Equations (3)–(15)). In the beginning, the SO 4   · radical was generated through the activation of PMS on activating sites (Co2+ and Mn3+) by Equations (3) and (4). The recycling of the redox pair of Co3+/Co2+ (1.8 V) could proceed as Equation (5) since it had higher standard redox potentials than that of SO 5   · / HSO 5 . Then, the Co3+ was reduced to Co2+ by the electrons from lattice oxygen. It needs to be noticed that E Mn 4 + / Mn 3 + 0 (0.15 V) was smaller than HSO 5 ( E HSO 5   · / SO 5 0 = 1.1 V), which shows the reduction was thermodynamically favorable (Equation (6)), and ·OH also exists through Equations (7)–(9), depending on the actual pH. Meanwhile, the standard redox potential of E Mn 4 + / Mn 3 + 0 (0.15 V) was lower, with a difference of 1.66V than E Co 3 + / Co 2 + 0 [64]. Therefore, electronic transfer tended to move from Mn3+ to Co3+ as Equation (10) shows, according to the thermodynamic favorable principles. It was validated by the relatively slight change in Mn valuation and sharp increase of Co3+ from the XPS spectra. In addition, the Mn4+ was reduced by the neighboring O L 2 as Equation (11) shows [28], which can explain the increase of O L and Mn3+. Then, the generated Mn3+ could participate in the redox process. Overall, the bimetallic synergistic effects may be seen as the crucial factor to activate PMS since the dopant of Mn cycling of Co-Mn was proven to be more efficient than the SrCoO3 and SrMnO3 perovskites.
Co 2 + + HSO 5   Co 3 + + SO 4   · + OH ,  
Mn 3 + + HSO 5   Mn 4 + + SO 4   · + OH
Co 3 + + HSO 5   Co 2 + + SO 5   · + H +
Mn 4 + + HSO 5   Mn 3 + + SO 5 ·   + H +
Co 2 + / Mn 3 + + HSO 5   Co 3 + / Mn 4 +   · OH + SO 4 2
SO 4   · + OH SO 4 2 + · OH  
SO 4   · + H 2 O   SO 4 2 + · OH + H +
Co 3 + + Mn 3 + Co 2 + + Mn 4 +   Δ E = 1.66   V  
Mn 4 + + O L 2   Mn 3 + + O L  
As a high electrophilic reactive oxygen species, singlet oxygen is gradually getting much more attention in recent research [65,66,67]. In general, singlet oxygen can derive from the decomposition of PMS in suitable pH (near pKa) [68] as Equation (12) shows. In this work, 1O2 could be generated through metal-based activators [69,70]. The oxygen vacancies could be converted into active oxygen(O*) and further changed into 1O2 (Equation (13)) with PMS. Meanwhile, the generated SO 5   · in the solution could react with H2O to obtain 1O2 (Equation (14)) also [71]. The SO 4   · , ·OH, and 1O2 radical generated in the SrCo0.5Mn0.5O3–PMS system acted together for RhB degradation. On the whole, the Mn species in the SrCo0.5Mn0.5O3 perovskite not only formed more oxygen vacancies but also participated in PMS activation as activate sites. The possible mechanism in SrCo0.5Mn0.5O3–PMS for removing RhB is depicted in Scheme 1.
HSO 5 + SO 5 2 HSO 4 + SO 4 2 + 1 O 2
HSO 5 + O *   HSO 4 + 1 O 2
2   SO 5   · + H 2 O     2   SO 4 2   + 2 H +   + 1.5   1 O 2
1 O 2 + SO 4   · + · OH + RhB     intermediates     CO 2 + H 2 O + SO 4 2

4. Conclusions

In this work, different SrCoxMn1−xO3 perovskites were synthesized to catalytically promote the degradation of organic contaminants by SR-AOPs. By adding the Mn element to a similar molar ratio of Co, SrCo0.5Mn0.5O3 still maintained the perovskite structure but has a comparatively higher specific surface area than SrCoO3, and it had more oxygen vacancy than other perovskites. Meanwhile, the SrCo0.5Mn0.5O3 perovskite also showed the highest performance for PMS activation compared to any other catalysts, and the promotion originated from the doping of Mn instead of the appearance of the SrMnO3 phase since the latter performed poorly in PMS activation. Moreover, the stability and reusability were found to perform well and SrCo0.5Mn0.5O3 still had a high activity for RhB removal after four cycles. The B sites composed of Co mixed with Mn were recognized to have a great effect on generating active radicals, which may be explained by two aspects: the generation of singlet oxygen by more oxygen vacancies and the redox cycling between Co and Mn. In the SrCo0.5Mn0.5O3–PMS system, SO 4   · , ·OH, and 1O2 were generated and worked together for the degradation of RhB. Our study prepared a high-performance catalyst for AOPs, which had a promising prospect for complex water treatment.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/pr11041279/s1. Table S1: Lattice parameters of SrCoO3 and SrCo0.5Mn0.5O3 samples. Table S2: the proportion of Co and Mn in fresh and used SrCo0.5Mn0.5O3 perovskite from XPS spectura. Table S3: the proportion of O species in fresh and used SrCo0.5Mn0.5O3 perovskite from XPS spectura. Figure S1: XRD patterens of SrCoxMn1−xO3 perovskites. Figure S2: Nitrogen adsorption-desorption isotherms of perovskite catalysts with different Mn substituted amounts. Figure S3: EDS mapping spectura of SrCo0.5Mn0.5O3 samples. Figure S4: FT-IR of SrCoxMn1−xO3 perovskites. Figure S5: Degradation efficience of RhB in the presence of SrCoO3 and PMS alone. Reaction conditions: PMS:0.06 mM, SrCoO3:0.03 g L−1, T = 298 k, pH = 7. Figure S6: pHpzc of SrCo0.5Mn0.5O3 perovskite. Figure S7: UV-VIS spectura of RHB in different time [72,73,74].

Author Contributions

Conceptualization, P.S., S.H., B.Z. and X.L.; Formal analysis, X.Y., C.Y., S.H., K.L., X.L., K.C., T.L. and J.J.; Investigation, X.Y., C.Y., X.L. and H.H.; Software, P.S., S.H. and B.Z.; Validation, P.S.; Writing—original draft, P.S., C.Y., S.H. and X.L.; Writing—review & editing, X.Y., B.Z., K.L., Z.Y. (Zhenyu Yang), Z.Y. (Zhiwei Yuan), Q.S. and J.R. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Natural Science Foundation of China (Grant No. 21673029) and the Science and Technology Project of Sinopec Group (No. H21002-3).

Data Availability Statement

The data that support the findings of this study are available from the authors, upon reasonable request.

Acknowledgments

We are grateful for the help of Institute of Engineering Technology of Sinopec Catalyst Co., Ltd.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Muñoz-García, A.B.; Benesperi, I.; Boschloo, G.; Concepcion, J.J.; Delcamp, J.H.; Gibson, E.A.; Meyer, G.J.; Pavone, M.; Pettersson, H.; Hagfeldt, A.; et al. Dye-sensitized solar cells strike back. Chem. Soc. Rev. 2021, 50, 12450–12550. [Google Scholar] [CrossRef]
  2. Dutta, S.; Gupta, B.; Srivastava, S.K.; Gupta, A.K. Recent advances on the removal of dyes from wastewater using various adsorbents: A critical review. Mater. Adv. 2021, 2, 4497–4531. [Google Scholar] [CrossRef]
  3. Ardila-Leal, L.D.; Poutou-Pinales, R.A.; Pedroza-Rodriguez, A.M.; Quevedo-Hidalgo, B.E. A Brief History of Colour, the Environmental Impact of Synthetic Dyes and Removal by Using Laccases. Molecules 2021, 26, 3813. [Google Scholar] [CrossRef] [PubMed]
  4. Ahmad, A.; Mohd-Setapar, S.H.; Chuong, C.S.; Khatoon, A.; Wani, W.A.; Kumar, R.; Rafatullah, M. Recent advances in new generation dye removal technologies: Novel search for approaches to reprocess wastewater. RSC Adv. 2015, 5, 30801–30818. [Google Scholar] [CrossRef]
  5. Rafaqat, S.; Ali, N.; Torres, C.; Rittmann, B. Recent progress in treatment of dyes wastewater using microbial-electro-Fenton technology. RSC Adv. 2022, 12, 17104–17137. [Google Scholar] [CrossRef] [PubMed]
  6. Muraro, P.C.L.; Mortari, S.R.; Vizzotto, B.S.; Chuy, G.; Dos Santos, C.; Brum, L.F.W.; da Silva, W.L. Iron oxide nanocatalyst with titanium and silver nanoparticles: Synthesis, characterization and photocatalytic activity on the degradation of Rhodamine B dye. Sci. Rep. 2020, 10, 3055. [Google Scholar] [CrossRef]
  7. Huang, D.; Zhang, G.; Yi, J.; Cheng, M.; Lai, C.; Xu, P.; Zhang, C.; Liu, Y.; Zhou, C.; Xue, W.; et al. Progress and challenges of metal-organic frameworks-based materials for SR-AOPs applications in water treatment. Chemosphere 2021, 263, 127672. [Google Scholar] [CrossRef]
  8. Gallego-Ramírez, C.; Chica, E.; Rubio-Clemente, A. Coupling of Advanced Oxidation Technologies and Biochar for the Removal of Dyes in Water. Water 2022, 14, 2531. [Google Scholar] [CrossRef]
  9. Liu, L.; Chen, Z.; Zhang, J.; Shan, D.; Wu, Y.; Bai, L.; Wang, B. Treatment of industrial dye wastewater and pharmaceutical residue wastewater by advanced oxidation processes and its combination with nanocatalysts: A review. J. Water Process Eng. 2021, 42, 102122. [Google Scholar] [CrossRef]
  10. Chen, Q.; Ji, F.; Guo, Q.; Guan, W.; Yan, P.; Pei, L.; Xu, X. Degradation of Phenol by Vis/Co-TiO2/KHSO5 Hybrid Co/SR–Photoprocess at Neutral pH. Ind. Eng. Chem. Res. 2013, 52, 12540–12549. [Google Scholar] [CrossRef]
  11. Lian, Q.; Roy, A.; Kizilkaya, O.; Gang, D.D.; Holmes, W.; Zappi, M.E.; Zhang, X.; Yao, H. Uniform Mesoporous Amorphous Cobalt-Inherent Silicon Oxide as a Highly Active Heterogeneous Catalyst in the Activation of Peroxymonosulfate for Rapid Oxidation of 2,4-Dichlorophenol: The Important Role of Inherent Cobalt in the Catalytic Mechanism. ACS Appl. Mater. Interfaces 2020, 12, 57190–57206. [Google Scholar] [CrossRef] [PubMed]
  12. Barndõk, H.; Hermosilla, D.; Negro, C.; Blanco, Á. Comparison and Predesign Cost Assessment of Different Advanced Oxidation Processes for the Treatment of 1,4-Dioxane-Containing Wastewater from the Chemical Industry. ACS Sustain. Chem. Eng. 2018, 6, 5888–5894. [Google Scholar] [CrossRef]
  13. Qu, J.; Dai, X.; Hu, H.-Y.; Huang, X.; Chen, Z.; Li, T.; Cao, Y.; Daigger, G.T. Emerging Trends and Prospects for Municipal Wastewater Management in China. ACS EST Eng. 2022, 2, 323–336. [Google Scholar] [CrossRef]
  14. 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]
  15. Li, Z.; Wang, M.; Jin, C.; Kang, J.; Liu, J.; Yang, H.; Zhang, Y.; Pu, Q.; Zhao, Y.; You, M.; et al. Synthesis of novel Co3O4 hierarchical porous nanosheets via corn stem and MOF-Co templates for efficient oxytetracycline degradation by peroxymonosulfate activation. Chem. Eng. J. 2020, 392, 123789. [Google Scholar] [CrossRef]
  16. Lee, Y.C.; Lo, S.L.; Kuo, J.; Huang, C.P. Promoted degradation of perfluorooctanic acid by persulfate when adding activated carbon. J. Hazard. Mater. 2013, 261, 463–469. [Google Scholar] [CrossRef]
  17. Antoniou, M.G.; de la Cruz, A.A.; Dionysiou, D.D. Degradation of microcystin-LR using sulfate radicals generated through photolysis, thermolysis and e− transfer mechanisms. Appl. Catal. B Environ. 2010, 96, 290–298. [Google Scholar] [CrossRef]
  18. Chen, M.; Zhu, L.; Liu, S.; Li, R.; Wang, N.; Tang, H. Efficient degradation of organic pollutants by low-level Co2+ catalyzed homogeneous activation of peroxymonosulfate. J. Hazard. Mater. 2019, 371, 456–462. [Google Scholar] [CrossRef]
  19. So, H.-L.; Lin, K.-Y.; Chu, W.; Gong, H. Degradation of Triclosan by Recyclable MnFe2O4-Activated PMS: Process Modification for Reduced Toxicity and Enhanced Performance. Ind. Eng. Chem. Res. 2020, 59, 4257–4264. [Google Scholar] [CrossRef]
  20. Yao, Y.; Yang, Z.; Zhang, D.; Peng, W.; Sun, H.; Wang, S. Magnetic CoFe2O4–Graphene Hybrids: Facile Synthesis, Characterization, and Catalytic Properties. Ind. Eng. Chem. Res. 2012, 51, 6044–6051. [Google Scholar] [CrossRef]
  21. Lin, N.; Gong, Y.; Wang, R.; Wang, Y.; Zhang, X. Critical review of perovskite-based materials in advanced oxidation system for wastewater treatment: Design, applications and mechanisms. J. Hazard. Mater. 2022, 424, 127637. [Google Scholar] [CrossRef] [PubMed]
  22. Ji, R.; Chen, J.; Liu, T.; Zhou, X.; Zhang, Y. Critical review of perovskites-based advanced oxidation processes for wastewater treatment: Operational parameters, reaction mechanisms, and prospects. Chin. Chem. Lett. 2022, 33, 643–652. [Google Scholar] [CrossRef]
  23. 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]
  24. Xu, X.; Wang, W.; Zhou, W.; Shao, Z. Recent Advances in Novel Nanostructuring Methods of Perovskite Electrocatalysts for Energy-Related Applications. Small Methods 2018, 2, 1800071. [Google Scholar] [CrossRef]
  25. Xu, X.; Pan, Y.; Ge, L.; Chen, Y.; Mao, X.; Guan, D.; Li, M.; Zhong, Y.; Hu, Z.; Peterson, V.K.; et al. High-Performance Perovskite Composite Electrocatalysts Enabled by Controllable Interface Engineering. Small 2021, 17, e2101573. [Google Scholar] [CrossRef]
  26. Hammouda, S.B.; 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]
  27. Pang, X.; Guo, Y.; Zhang, Y.; Xu, B.; Qi, F. LaCoO3 perovskite oxide activation of peroxymonosulfate for aqueous 2-phenyl-5-sulfobenzimidazole degradation: Effect of synthetic method and the reaction mechanism. Chem. Eng. J. 2016, 304, 897–907. [Google Scholar] [CrossRef]
  28. 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]
  29. Jing, J.; Pervez, M.N.; Sun, P.; Cao, C.; Li, B.; Naddeo, V.; Jin, W.; Zhao, Y. Highly efficient removal of bisphenol A by a novel Co-doped LaFeO3 perovskite/PMS system in salinity water. Sci. Total Environ. 2021, 801, 149490. [Google Scholar] [CrossRef]
  30. Koo, P.-L.; Jaafar, N.F.; Yap, P.-S.; Oh, W.-D. A review on the application of perovskite as peroxymonosulfate activator for organic pollutants removal. J. Environ. Chem. Eng. 2022, 10, 107093. [Google Scholar] [CrossRef]
  31. Najjar, H.; Batis, H. Development of Mn-based perovskite materials: Chemical structure and applications. Catal. Rev. 2016, 58, 371–438. [Google Scholar] [CrossRef]
  32. Wang, T.; Qian, X.; Yue, D.; Yan, X.; Yamashita, H.; Zhao, Y. CaMnO3 perovskite nanocrystals for efficient peroxydisulfate activation. Chem. Eng. J. 2020, 398, 125638. [Google Scholar] [CrossRef]
  33. 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] [PubMed]
  34. Pan, Y.; Xu, X.; Zhong, Y.; Ge, L.; Chen, Y.; Veder, J.M.; Guan, D.; O’Hayre, R.; Li, M.; Wang, G.; et al. Direct evidence of boosted oxygen evolution over perovskite by enhanced lattice oxygen participation. Nat. Commun. 2020, 11, 2002. [Google Scholar] [CrossRef] [PubMed]
  35. Jeerh, G.; Zou, P.; Zhang, M.; Tao, S. Perovskite oxide LaCr0.25Fe0.25Co0.5O3−δ as an efficient non-noble cathode for direct ammonia fuel cells. Appl. Catal. B: Environ. 2022, 319, 121919. [Google Scholar] [CrossRef]
  36. Wang, G.; Cheng, C.; Zhu, J.; Wang, L.; Gao, S.; Xia, X. Enhanced degradation of atrazine by nanoscale LaFe1−xCuxO3-delta perovskite activated peroxymonosulfate: Performance and mechanism. Sci. Total Environ. 2019, 673, 565–575. [Google Scholar] [CrossRef] [PubMed]
  37. Yang, L.; Jiao, Y.; Xu, X.; Pan, Y.; Su, C.; Duan, X.; Sun, H.; Liu, S.; Wang, S.; Shao, Z. Superstructures with Atomic-Level Arranged Perovskite and Oxide Layers for Advanced Oxidation with an Enhanced Non-Free Radical Pathway. ACS Sustain. Chem. Eng. 2022, 10, 1899–1909. [Google Scholar] [CrossRef]
  38. He, L.; Zhang, Y. Singlet oxygen produced SrCoO2.5 in environmental protection: Extraordinary electronic properties and promoted catalytic performance. J. Sol-Gel Sci. Technol. 2021, 99, 391–401. [Google Scholar] [CrossRef]
  39. Zhang, H.; Zhang, R.; Wu, Z.; Yang, F.; Luo, M.; Yao, G.; Ao, Z.; Lai, B. Cobalt-doped boosted the peroxymonosulfate activation performance of LaFeO3 perovskite for atrazine degradation. Chem. Eng. J. 2023, 452, 139427. [Google Scholar] [CrossRef]
  40. Chen, H.; Lim, C.; Zhou, M.; He, Z.; Sun, X.; Li, X.; Ye, Y.; Tan, T.; Zhang, H.; Yang, C.; et al. Activating Lattice Oxygen in Perovskite Oxide by B-Site Cation Doping for Modulated Stability and Activity at Elevated Temperatures. Adv. Sci. 2021, 8, e2102713. [Google Scholar] [CrossRef]
  41. Khazaei, M.; Malekzadeh, A.; Amini, F.; Mortazavi, Y.; Khodadadi, A. Effect of citric acid concentration as emulsifier on perovskite phase formation of nano-sized SrMnO3 and SrCoO3 samples. Cryst. Res. Technol. 2010, 45, 1064–1068. [Google Scholar] [CrossRef]
  42. Giroir-Fendler, A.; Alves-Fortunato, M.; Richard, M.; Wang, C.; Díaz, J.A.; Gil, S.; Zhang, C.; Can, F.; Bion, N.; Guo, Y. Synthesis of oxide supported LaMnO3 perovskites to enhance yields in toluene combustion. Appl. Catal. B Environ. 2016, 180, 29–37. [Google Scholar] [CrossRef]
  43. Zhao, L.; Han, T.; Wang, H.; Zhang, L.; Liu, Y. Ni-Co alloy catalyst from LaNi1−xCoxO3 perovskite supported on zirconia for steam reforming of ethanol. Appl. Catal. B: Environ. 2016, 187, 19–29. [Google Scholar] [CrossRef]
  44. Mao, W.; Fan, Y.; Hu, X. Degradation of tetrabromobisphenol A through peroxymonaosulfate oxidation activated by La0.5Sr0.5CoxMn1−xO3-delta perovskite. Environ. Sci. Pollut. Res. Int. 2021, 28, 65814–65821. [Google Scholar] [CrossRef]
  45. Yang, L.; Hu, R.; Li, H.; Jia, Y.; Zhou, Q.; Wang, H. The effect of interaction between La2AlCoO6 and CuCl2 on ethane oxychlorination. J. Ind. Eng. Chem. 2017, 56, 120–128. [Google Scholar] [CrossRef]
  46. Roozbahani, H.; Maghsoodi, S.; Raei, B.; Kootenaei, A.S.; Azizi, Z. Effects of catalyst preparation methods on the performance of La2MMnO6 (M=Co, Ni) double perovskites in catalytic combustion of propane. Korean J. Chem. Eng. 2022, 39, 586–595. [Google Scholar] [CrossRef]
  47. Li, R.; Zhang, L.; Zhu, S.; Fu, S.; Dong, X.; Ida, S.; Zhang, L.; Guo, L. Layered δ-MnO2 as an active catalyst for toluene catalytic combustion. Appl. Catal. A Gen. 2020, 602, 117715. [Google Scholar] [CrossRef]
  48. Chen, X.; Zhou, J.; Yang, H.; Wang, H.; Li, H.; Wu, S.; Yang, W. PMS activation by magnetic cobalt-N-doped carbon composite for ultra-efficient degradation of refractory organic pollutant: Mechanisms and identification of intermediates. Chemosphere 2022, 287, 132074. [Google Scholar] [CrossRef]
  49. 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. 2019, 59, 99–109. [Google Scholar] [CrossRef]
  50. Zhang, W.; Su, Y.; Zhang, X.; Yang, Y.; Guo, X. Facile synthesis of porous NiCo2O4 nanoflakes as magnetic recoverable catalysts towards the efficient degradation of RhB. RSC Adv. 2016, 6, 64626–64633. [Google Scholar] [CrossRef]
  51. Yu, H.; Ding, D.; Zhao, S.; Faheem, M.; Mao, W.; Yang, L.; Chen, L.; Cai, T. Co/N co-doped porous carbon as a catalyst for the degradation of RhB by efficient activation of peroxymonosulfate. Environ. Sci. Pollut. Res. 2022, 30, 10969–10981. [Google Scholar] [CrossRef] [PubMed]
  52. Xu, Z.; Wu, Y.; Ji, Q.; Li, T.; Xu, C.; Qi, C.; He, H.; Yang, S.; Li, S.; Yan, S.; et al. Understanding spatial effects of tetrahedral and octahedral cobalt cations on peroxymonosulfate activation for efficient pollution degradation. Appl. Catal. B Environ. 2021, 291, 120072. [Google Scholar] [CrossRef]
  53. Tan, J.; Xu, C.; Zhang, X. MOFs-derived defect carbon encapsulated magnetic metallic Co nanoparticles capable of efficiently activating PMS to rapidly degrade dyes. Sep. Purif. Technol. 2022, 289, 120812. [Google Scholar] [CrossRef]
  54. Wang, L.; Di, J.; Nie, J.; Ma, G. Multicomponent Doped Sugar-Coated Nanofibers for Peroxymonosulfate Activation. ACS Appl. Nano Mater. 2019, 2, 6998–7007. [Google Scholar] [CrossRef]
  55. Liu, Y.; Tian, X.; Xiong, W.; Nie, G.; Xiao, L. Prussian blue analogs derived nanostructured Mn/Fe bimetallic carbon materials for organic pollutants degradation via peroxymonosulfate activation. Colloids Surf. A Physicochem. Eng. Asp. 2023, 657, 130592. [Google Scholar] [CrossRef]
  56. Fan, J.; Zhao, Z.; Ding, Z.; Liu, J. Synthesis of different crystallographic FeOOH catalysts for peroxymonosulfate activation towards organic matter degradation. RSC Adv. 2018, 8, 7269–7279. [Google Scholar] [CrossRef]
  57. Yang, Z.; Duan, X.; Wang, J.; Li, Y.; Fan, X.; Zhang, F.; Zhang, G.; Peng, W. Facile Synthesis of High-Performance Nitrogen-Doped Hierarchically Porous Carbon for Catalytic Oxidation. ACS Sustain. Chem. Eng. 2020, 8, 4236–4243. [Google Scholar] [CrossRef]
  58. Zhang, H.; Zhou, C.; Zeng, H.; Shi, Z.; Wu, H.; Deng, L. Novel sulfur vacancies featured MIL-88A(Fe)@CuS rods activated peroxymonosulfate for coumarin degradation: Different reactive oxygen species generation routes under acidic and alkaline pH. Process Saf. Environ. Prot. 2022, 166, 11–22. [Google Scholar] [CrossRef]
  59. Liang, P.; Zhang, C.; Duan, X.; Sun, H.; Liu, S.; Tade, M.O.; Wang, S. N-Doped Graphene from Metal–Organic Frameworks for Catalytic Oxidation of p-Hydroxylbenzoic Acid: N-Functionality and Mechanism. ACS Sustain. Chem. Eng. 2017, 5, 2693–2701. [Google Scholar] [CrossRef]
  60. Zhu, M.; Miao, J.; Duan, X.; Guan, D.; Zhong, Y.; Wang, S.; Zhou, W.; Shao, Z. Postsynthesis Growth of CoOOH Nanostructure on SrCo0.6Ti0.4O3−δ Perovskite Surface for Enhanced Degradation of Aqueous Organic Contaminants. ACS Sustain. Chem. Eng. 2018, 6, 15737–15748. [Google Scholar] [CrossRef]
  61. Wang, L.; Fu, Y.; Li, Q.; Wang, Z. EPR Evidence for Mechanistic Diversity of Cu(II)/Peroxygen Oxidation Systems by Tracing the Origin of DMPO Spin Adducts. Environ. Sci. Technol. 2022, 56, 8796–8806. [Google Scholar] [CrossRef] [PubMed]
  62. Huang, B.C.; Huang, G.X.; Jiang, J.; Liu, W.J.; Yu, H.Q. Carbon-Based Catalyst Synthesized and Immobilized under Calcium Salt Assistance To Boost Singlet Oxygen Evolution for Pollutant Degradation. ACS Appl. Mater. Interfaces 2019, 11, 43180–43187. [Google Scholar] [CrossRef] [PubMed]
  63. Yu, D.; He, J.; Xie, T.; Yang, J.; Wang, J.; Xie, J.; Shi, H.; Gao, Z.; Xiang, B.; Dionysiou, D.D. Boosting catalytic activity of SrCoO2.52 perovskite by Mn atom implantation for advanced peroxymonosulfate activation. J. Hazard. Mater. 2023, 442, 130085. [Google Scholar] [CrossRef]
  64. Xu, Y.; Hu, E.; Xu, D.; Guo, Q. Activation of peroxymonosulfate by bimetallic CoMn oxides loaded on coal fly ash-derived SBA-15 for efficient degradation of Rhodamine B. Sep. Purif. Technol. 2021, 274, 119081. [Google Scholar] [CrossRef]
  65. 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]
  66. Li, X.; Wang, L.; Guo, Y.; Song, W.; Li, Y.; Yan, L. Goethite-MoS2 hybrid with dual active sites boosted peroxymonosulfate activation for removal of tetracycline: The vital roles of hydroxyl radicals and singlet oxygen. Chem. Eng. J. 2022, 450, 138104. [Google Scholar] [CrossRef]
  67. Shao, S.; Li, X.; Gong, Z.; Fan, B.; Hu, J.; Peng, J.; Lu, K.; Gao, S. A new insight into the mechanism in Fe3O4@CuO/PMS system with low oxidant dosage. Chem. Eng. J. 2022, 438, 135474. [Google Scholar] [CrossRef]
  68. Zhou, Y.; Jiang, J.; Gao, Y.; Ma, J.; Pang, S.Y.; Li, J.; Lu, X.T.; Yuan, L.P. Activation of Peroxymonosulfate by Benzoquinone: A Novel Nonradical Oxidation Process. Environ. Sci. Technol. 2015, 49, 12941–12950. [Google Scholar] [CrossRef]
  69. Zhu, S.; Li, X.; Kang, J.; Duan, X.; Wang, S. Persulfate Activation on Crystallographic Manganese Oxides: Mechanism of Singlet Oxygen Evolution for Nonradical Selective Degradation of Aqueous Contaminants. Environ. Sci. Technol. 2019, 53, 307–315. [Google Scholar] [CrossRef]
  70. Bu, Y.; Li, H.; Yu, W.; Pan, Y.; Li, L.; Wang, Y.; Pu, L.; Ding, J.; Gao, G.; Pan, B. Peroxydisulfate Activation and Singlet Oxygen Generation by Oxygen Vacancy for Degradation of Contaminants. Environ. Sci. Technol. 2021, 55, 2110–2120. [Google Scholar] [CrossRef]
  71. Yang, S.; Zhang, S.; Li, X.; Du, Y.; Xing, Y.; Xu, Q.; Wang, Z.; Li, L.; Zhu, X. One-step pyrolysis for the preparation of sulfur-doped biochar loaded with iron nanoparticles as an effective peroxymonosulfate activator for RhB degradation. New J. Chem. 2022, 46, 5678–5689. [Google Scholar] [CrossRef]
  72. Zhao, P.; Fang, F.; Feng, N.; Chen, C.; Liu, G.; Chen, L.; Zhu, Z.; Meng, J.; Wan, H.; Guan, G. Self-templating construction of mesopores on three dimensionally ordered macroporous La0.5Sr0.5MnO3 perovskite with enhanced performance for soot combustion. Catal. Sci. Technol. 2019, 9, 1835–1846. [Google Scholar] [CrossRef]
  73. Zhao, S.; Wang, L.; Wang, Y.; Li, X. Hierarchically porous LaFeO3 perovskite prepared from the pomelo peel bio-template for catalytic oxidation of NO. J. Phys. Chem. Solids 2018, 116, 43–49. [Google Scholar] [CrossRef]
  74. Demirel, S.; Oz, E.; Altin, S.; Bayri, A.; Baglayan, O.; Altin, E.; Avci, S. Structural, magnetic, electrical and electrochemical properties of SrCoO2.5, Sr9Co2Mn5O21 and SrMnO3 compounds. Ceram. Int. 2017, 43, 14818–14826. [Google Scholar] [CrossRef]
Figure 1. (a) XRD patterns of the SrCoxMn1−xO3 perovskites. (b) A comparison with the pattern of the SrCoxMn1−xO3 perovskites.
Figure 1. (a) XRD patterns of the SrCoxMn1−xO3 perovskites. (b) A comparison with the pattern of the SrCoxMn1−xO3 perovskites.
Processes 11 01279 g001
Figure 2. The 1μm and 5μm SEM images of SrCoxMn1−xO3. (a,d) SrCoO3, (b,e) SrCo0.5Mn0.5O3, (c,f) SrMnO3 perovskites; (g) EDS spectra of SrCo0.5Mn0.5O3.
Figure 2. The 1μm and 5μm SEM images of SrCoxMn1−xO3. (a,d) SrCoO3, (b,e) SrCo0.5Mn0.5O3, (c,f) SrMnO3 perovskites; (g) EDS spectra of SrCo0.5Mn0.5O3.
Processes 11 01279 g002
Figure 3. (a) O2-TPD of SrCoxMn1-xO3 perovskites. (b) EPR spectra of SrCoxMn1−xO3 perovskites.
Figure 3. (a) O2-TPD of SrCoxMn1-xO3 perovskites. (b) EPR spectra of SrCoxMn1−xO3 perovskites.
Processes 11 01279 g003
Figure 4. (a) Degradation efficiency of RhB in the presence of SrCoxMn1−xO3 perovskites. Reaction conditions: PMS:0.06mM, catalysts:0.06g L−1, 298 K, pH = 7. (b) Degradation efficiency of RhB in the presence of SrCoxMn1−xO3 perovskites. (c) Leaching concentration of Co and Mn of SrCoxMn1−xO3 perovskites. (d) The degradation efficiency of RhB in the Co2+–PMS system.
Figure 4. (a) Degradation efficiency of RhB in the presence of SrCoxMn1−xO3 perovskites. Reaction conditions: PMS:0.06mM, catalysts:0.06g L−1, 298 K, pH = 7. (b) Degradation efficiency of RhB in the presence of SrCoxMn1−xO3 perovskites. (c) Leaching concentration of Co and Mn of SrCoxMn1−xO3 perovskites. (d) The degradation efficiency of RhB in the Co2+–PMS system.
Processes 11 01279 g004
Figure 5. Reusability of SrCo0.5Mn0.5O3 for the degradation of RhB. Reaction conditions: [RhB] = 40 mg L−1, [PMS] = 0.06 mM, catalysts = 0.03 g L−1, T = 298 K.
Figure 5. Reusability of SrCo0.5Mn0.5O3 for the degradation of RhB. Reaction conditions: [RhB] = 40 mg L−1, [PMS] = 0.06 mM, catalysts = 0.03 g L−1, T = 298 K.
Processes 11 01279 g005
Figure 6. Effects of (a) catalyst dosage, (b) PMS concentration, and (c) pH on RhB degradation in SrCo0.5Mn0.5O3−PMS (PMS: 0.04 mM). Reaction conditions: RhB: 40 mg L−1, PMS: 0.06 mM, SrCo0.5Mn0.5O3: 0.03 g L−1, unless otherwise noted. T: 298.15 K. pH: nature.
Figure 6. Effects of (a) catalyst dosage, (b) PMS concentration, and (c) pH on RhB degradation in SrCo0.5Mn0.5O3−PMS (PMS: 0.04 mM). Reaction conditions: RhB: 40 mg L−1, PMS: 0.06 mM, SrCo0.5Mn0.5O3: 0.03 g L−1, unless otherwise noted. T: 298.15 K. pH: nature.
Processes 11 01279 g006
Figure 7. (a) Quenching tests in the SrCo0.5Mn0.5O3–PMS system. (b) EPR spectra of DMPO-radical adducts and (c) singlet oxygen in the SrCo0.5Mn0.5O3–PMS reaction systems.
Figure 7. (a) Quenching tests in the SrCo0.5Mn0.5O3–PMS system. (b) EPR spectra of DMPO-radical adducts and (c) singlet oxygen in the SrCo0.5Mn0.5O3–PMS reaction systems.
Processes 11 01279 g007
Figure 8. (a) survey XPS spectra of fresh and used SrCo0.5Mn0.5O3 perovskite. (bd) XPS spectra of Co 2p, Mn 2p O 1s of fresh and used SrCo0.5Mn0.5O3 perovskites.
Figure 8. (a) survey XPS spectra of fresh and used SrCo0.5Mn0.5O3 perovskite. (bd) XPS spectra of Co 2p, Mn 2p O 1s of fresh and used SrCo0.5Mn0.5O3 perovskites.
Processes 11 01279 g008
Scheme 1. A possible mechanism for activation of PMS by SrCo0.5Mn0.5O3.
Scheme 1. A possible mechanism for activation of PMS by SrCo0.5Mn0.5O3.
Processes 11 01279 sch001
Table 1. The actual composition of SrCoxMn1−xO3 perovskites.
Table 1. The actual composition of SrCoxMn1−xO3 perovskites.
PerovskitesMetal Content (Atomic)
CoMn
SrCo0.7Mn0.3O30.710.29
SrCo0.5Mn0.5O30.520.48
SrCo0.3Mn0.7O30.330.67
Table 2. BET surface area, total pore volume, and average pore diameter of the SrCoxMn1−xO3 perovskites.
Table 2. BET surface area, total pore volume, and average pore diameter of the SrCoxMn1−xO3 perovskites.
PerovskitesSSR (m2/g)Pore Volume (cm3/g)Average Pore Size (nm)
SrCoO30.630.00212.2
SrCo0.7Mn0.3O33.050.01113.9
SrCo0.5Mn0.5O33.330.00911.9
SrCo0.3Mn0.7O34.210.01617.2
SrMnO34.110.01617.2
Table 3. Summary of different catalysts–PMS systems for removing RhB.
Table 3. Summary of different catalysts–PMS systems for removing RhB.
CatalystsConditionsDegradationMain Active SpeciesIon Leaching/mg L−1Ref.
PMS Conc/mMRhB Conc/mg L−1Catalyst Dose/g L−1
Co-NC-8500.80080.000.02520 min 100%Co25[48]
LMO-5bar-6004.90020.000.20045 min 100%Mn0.200[49]
NiCo2O40.50023.950.20020 min 95%Co Ni-[50]
Co/N-NPC0.5/9001.2509.580.00510 min 100%Co0.408[51]
ZnCo2O40.04910.000.03030 min 100%Co0.220[52]
Co@DC0.16010.000.0205 min 100%Co0.840[53]
Fe/Co-N/P-90.65040.000.06035 min 98.31%C-π, Co, Fe.-[54]
Mn-Fe-CN0.32025.000.10012 min 100%Fe0.600[55]
SrCo0.5Mn0.5O30.06040.000.06015 min 100%Co Mn0.220This work
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Shao, P.; Yin, X.; Yu, C.; Han, S.; Zhao, B.; Li, K.; Li, X.; Yang, Z.; Yuan, Z.; Shi, Q.; et al. Enhanced Activation of Peroxymonosulfate via Sulfate Radicals and Singlet Oxygen by SrCoxMn1−xO3 Perovskites for the Degradation of Rhodamine B. Processes 2023, 11, 1279. https://doi.org/10.3390/pr11041279

AMA Style

Shao P, Yin X, Yu C, Han S, Zhao B, Li K, Li X, Yang Z, Yuan Z, Shi Q, et al. Enhanced Activation of Peroxymonosulfate via Sulfate Radicals and Singlet Oxygen by SrCoxMn1−xO3 Perovskites for the Degradation of Rhodamine B. Processes. 2023; 11(4):1279. https://doi.org/10.3390/pr11041279

Chicago/Turabian Style

Shao, Penghui, Xiping Yin, Chenyu Yu, Shuai Han, Baohuai Zhao, Kezhi Li, Xiang Li, Zhenyu Yang, Zhiwei Yuan, Qinzhi Shi, and et al. 2023. "Enhanced Activation of Peroxymonosulfate via Sulfate Radicals and Singlet Oxygen by SrCoxMn1−xO3 Perovskites for the Degradation of Rhodamine B" Processes 11, no. 4: 1279. https://doi.org/10.3390/pr11041279

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