Next Article in Journal
Synthesis, In Vitro, and In Silico Analysis of the Antioxidative Activity of Dapsone Imine Derivatives
Previous Article in Journal
Intermolecular Forces Driving Hexamethylenetetramine Co-Crystal Formation, a DFT and XRD Analysis
Previous Article in Special Issue
H2S Removal from Groundwater by Chemical Free Advanced Oxidation Process Using UV-C/VUV Radiation
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

A Review of Manganese(III) (Oxyhydr)Oxides Use in Advanced Oxidation Processes

1
Institut de Chimie de Clermont-Ferrand, Université Clermont Auvergne, CNRS, Clermont Auvergne INP SIGMA Clermont, F-63000 Clermont-Ferrand, France
2
École Nationale Supérieure de Chimie de Rennes, Université Rennes, CNRS, ISCR–UMR6226, F-35000 Rennes, France
3
Institut Universitaire de France (IUF), MESRI, 1 rue Descartes, 75231 Paris, France
*
Author to whom correspondence should be addressed.
Molecules 2021, 26(19), 5748; https://doi.org/10.3390/molecules26195748
Submission received: 3 September 2021 / Revised: 20 September 2021 / Accepted: 21 September 2021 / Published: 22 September 2021
(This article belongs to the Special Issue Applications of Advanced Oxidation Processes for Water Treatment)

Abstract

:
The key role of trivalent manganese (Mn(III)) species in promoting sulfate radical-based advanced oxidation processes (SR-AOPs) has recently attracted increasing attention. This review provides a comprehensive summary of Mn(III) (oxyhydr)oxide-based catalysts used to activate peroxymonosulfate (PMS) and peroxydisulfate (PDS) in water. The crystal structures of different Mn(III) (oxyhydr)oxides (such as α-Mn2O3, γ-MnOOH, and Mn3O4) are first introduced. Then the impact of the catalyst structure and composition on the activation mechanisms are discussed, as well as the effects of solution pH and inorganic ions. In the Mn(III) (oxyhydr)oxide activated SR-AOPs systems, the activation mechanisms of PMS and PDS are different. For example, both radical (such as sulfate and hydroxyl radical) and non-radical (singlet oxygen) were generated by Mn(III) (oxyhydr)oxide activated PMS. In comparison, the activation of PDS by α-Mn2O3 and γ-MnOOH preferred to form the singlet oxygen and catalyst surface activated complex to remove the organic pollutants. Finally, research gaps are discussed to suggest future directions in context of applying radical-based advanced oxidation in wastewater treatment processes.

1. Introduction

Over the past few decades, with the rapid development of industrialization and the increase of anthropogenic activities, huge amounts of organic and inorganic contaminants were discharged into the surface and ground waters, causing water pollution problems and threatening human health [1,2,3]. However, conventional water treatment technologies, such as filtration [4,5], precipitation [6,7], coagulation–flocculation [8,9,10], and biological treatment [11,12] exhibited a minimal effect on the removal of recalcitrant pollutants. Therefore, there is an increasing demand for efficient, economical, and environmental-friendly water treatment technologies. Advanced oxidation processes (AOPs) have attracted particular attention due to their high efficiency for removal of recalcitrant contaminant. AOPs are able to remove and mineralize most unbiodegradable pollutants into harmless compounds, such as CO2, H 2 O , and inorganic ions [13]. Based on various reaction conditions, AOPs can be classified into different categories, including Fenton reaction [14], Fenton-like reaction [15,16], photochemical oxidation [17,18], ultrasonic oxidation [19,20], electrochemical oxidation [21,22], ozone oxidation [23,24], and sulfate radical-based AOPs (SR-AOPs) [25,26,27]. Among them, the application of SR-AOPs for the removal of stubborn pollutants has received increasing attention due to their advantages. For instance, sulfate radical ( SO 4 ) has a longer lifetime compared with the hydroxyl radical ( HO ), a wide range of pH adaptation, and a high reduction potential (2.5-3.1 V vs. NHE) [28].
Generally, the peroxydisulfate (PDS, S 2 O 8 2 ) and peroxymonosulfate anions (PMS, HSO 5 ) are employed as the radical precursors for producing sulfate radicals through breaking the O-O bonds of precursors. In comparison with PMS, PDS has a longer O-O bonds distance (1.497 vs. 1.460 Å) and lower bond energy (140 vs. 140–213.3 kJ/mol) [29,30]. Therefore, PDS is theoretically easier than PMS to be cleaved to generate SO 4 . However, considering the unsymmetrical structure of PMS, it was reported that PMS activation was convenient for the removal of organic pollutants [31,32]. There are various ways to activate PMS and PDS to produce SO 4 , for example, heat, UV, alkaline solution, metal ions, and minerals [33,34,35,36,37].
The activation of PMS/PDS by different transition metal ions (i.e., Co(II), Ru(III), Fe(II), Fe(III), Ag(I), Mn(II), Ni(I), and V(III)) for organic pollutant degradation has been reported [32]. The results showed that PMS can be efficiently activated by Co(II) and Ru(III), while Ag(I) was identified as the best catalyst for PDS activation. However, the high price of Ag(I), Ru(III), and Co(II) restricts their application in practical water treatment. In comparison, the activation of PMS/PDS by the transition metal-based minerals (such as magnetite, birnessite, and manganite) has attracted much attention due to their various advantages, such as wide resources, easy recycling, and low energy requirement [38,39]. Among the transition metal oxides, the manganese oxides have been widely developed in PMS/PDS activation for recalcitrant pollutant degradation due to their excellent properties, such as various Mn valences, ubiquitous existence, cost-efficiency, and low toxicity [40]. For instance, Zhu et al. employed the β-MnO2 nanorods to activate PDS for the removal of phenol. Efficient degradation of phenol was achieved in β-MnO2/PDS system through the generation of singlet oxygen (1 O 2 ) [41]. Zhou et al. indicated the higher catalytic property of α-MnO2 than δ-MnO2 in PMS activation for 4-nitrophenol degradation because α-MnO2 owns more active sites, larger Brunauer–Emmett–Teller (BET) area, faster electron transfer rate, and better adsorption performance [42]. Furthermore, the activation of PMS by MnO2 with different crystal phases (i.e., α-, β-, γ-, and δ-MnO2) was reported by Huang et al. [43]. The results demonstrated the important role of crystalline structure and Mn(III) content on the catalytic reactivity of MnO2. Saputra et al. investigated the effect of Mn oxidation states (such as MnO, Mn2O3, Mn3O4, and MnO2) on the activation of PMS for phenol degradation. The results showed that Mn2O3 has the highest ability on PMS activation among these four manganese oxides [44]. Therefore, the structure of manganese oxides and the content of Mn(III) on the surface of manganese oxides play a critical role in the oxidative and catalytic reactivity of manganese oxides. The performance of MnO2 on PDS/PMS activation was well summarized in previous reviews [45,46,47]. However, no attempt has been made to provide a comprehensive review on Mn(III) (oxyhydr)oxides activated PMS/PDS for recalcitrant pollutants removal.
In light of the above information, this review aims to provide a comprehensive summary of reported Mn(III)-based catalysts in activating PMS/PDS. The structures of commonly used Mn(III) (oxyhydr)oxides (α-Mn2O3, Mn3O4, and γ-MnOOH) are first presented, then the effect of structure on the reactivity of Mn(III) (oxyhydr)oxides are discussed. Moreover, the radical and non-radical mechanisms of PMS/PDS activation by a single or combined Mn(III) species are summarized and the influence factors affecting the reactivity of Mn(III) (oxyhydr)oxides are introduced.
We are convinced that this review article will be of significant interest for researchers working on chemical oxidation for water decontamination processes. Finally, we also highlight how the literature lacks information and data that are crucial prior to high-scale applications.

2. Effect of Structure on the Reactivity of Mn(III) (Oxyhydr)Oxides

The oxidative and catalytic performance of manganese oxides can be affected by various structural factors including crystal phases, morphologies, crystal facets, and structural dimensionalities [48]. For instance, Huang et al. reported that δ -MnO2 showed higher oxidative activity than α - ,   β - , γ - ,   λ - MnO2 on bisphenol A oxidation due to the occurrence of more accessible active sites in layered δ -MnO2 than other tunnel structured MnO2 [49]. The authors also demonstrated the effects of structured MnO2 on peroxymonosulfate (PMS) activation, and the low reactivity of δ-MnO2 was attributed to its less crystallinity [43].
Crystalline manganese oxides are generally built on the same basic unit [MnO6] octahedral with the edges or corners sharing [41]. The commonly reported Mn(III) (oxyhydr)oxides include manganese(III) oxide (α-Mn2O3), groutite (α-MnOOH), feitknechtite (ꞵ-MnOOH), manganite (γ-MnOOH), and hausmannite (Mn3O4). The structures covered in the name of Mn(III) (oxyhydr)oxides are summarized in Table 1. Among them, α-Mn2O3, γ-MnOOH, and Mn3O4 have attracted increasing attention from the scientific community because of their promising technological applications, such as in catalysis, water treatment, and ion exchange. The crystalline structure of α-Mn2O3 was recognized as the body-centered cubic bixbyite phase, as shown in Figure 1a. γ-MnOOH possesses a typical (1 × 1) tunnel structure constructed by [MnO6] octahedral sharing the corners (Figure 1b). The structure of γ-MnOOH is analogous to that of pyrolusite, except that one-half of the oxygen atoms are replaced by hydroxyl anions compared with pyrolusite. For the crystalline Mn3O4, it exhibits a normal spinel structure with the formula Mn2+(Mn3+)2O4 where the Mn2+ and Mn3+ ions occupy the tetrahedral and octahedral sites, respectively (Figure 1c).
The influence of structures in the reactivity of common Mn(III) (oxyhydr)oxides is summarized in Table 2. For instance, Saputra et al. investigated the effect of morphology on the oxidation of phenol by Mn2O3 activated PMS. The results showed that cubic-Mn2O3 has the highest reactivity on PMS activation in comparison with octahedral- and truncated octahedral-Mn2O3, and it was due to the high surface area and distinct surface atoms arrangement of cubic-Mn2O3 [55]. Similarly, Cheng et al. successfully prepared three α-Mn2O3 in cubic-, truncated octahedral-, and octahedral-structure, and investigated the effect of crystal facets on the combustion of soot [56]. The results show that the soot combustion efficiency followed the order of α-Mn2O3-cubic > α-Mn2O3-truncated octahedral > α-Mn2O3-octahedral. The enhanced reactivity of α-Mn2O3-cubic was explained by the fact that the exposed (001) surface facets of α-Mn2O3-cubic have higher amounts of low-coordinated surface oxygen sites, which are capable of facilitating the oxygen activation and improving the surface redox properties.
In addition to α-Mn2O3, it was also reported that the oxidative and catalytic performances of Mn3O4 and γ-MnOOH were affected by their structures. For example, Ji et al. reported that the hexagonal nanoplate Mn3O4 exhibited superior catalytic performance on diesel soot combustion compared to the octahedral and nanoparticle Mn3O4, and the finding was explained by the improved amount of surface Mn4+ species and surface reactive oxygen species due to the increased fraction of exposed (112) facets in hexagonal nanoplate Mn3O4 [57]. The effect of morphology was also discovered by Liu et al., which demonstrated that the nanoflake Mn3O4 (exposure of (001) facet) has the highest oxygen reduction reactivity in comparison to nanoparticle Mn3O4 and nanorod Mn3O4 (exposure of (101) facet) [58]. In addition, He et al. investigated the activation of PMS by γ-MnOOH with different shapes, and the results showed that the catalytic activity of γ-MnOOH followed the order of nanowires > multi-branches > nanorods [59]. Different physicochemical parameters, such as specific surface area, Lewis sites, zeta-potential, and redox potential were measured to study the reason for the different catalytic performances of γ-MnOOH with distinct morphologies. It was found that the charge density on the surface played a crucial role in the interfacial reactivity between PMS and γ-MnOOH. In summary, the reactivity of Mn(III) (oxyhydr)oxides on radical precursor activation and pollutant oxidation can be deeply affected by their structures. The desirable morphologies and facets (such as cubic structure with (001) facet exposure) can apparently improve the reactivity of Mn(III) (oxyhydr)oxides.
Table 2. The effect of structures on the reactivity of Mn(III) (oxyhydr)oxides.
Table 2. The effect of structures on the reactivity of Mn(III) (oxyhydr)oxides.
CatalystsStructureInitial ConditionsReactivityMechanismRef. 1
α-Mn2O3Cubic;
octahedral;
truncated octahedral
[Catalyst] = 0.4 g/L;
[PMS] = 2 g/L;
[Phenol] = 25 ppm;
100% of phenol removal by cubic-Mn2O3 in 60 minHigh surface area and surface atoms arrangement of cubic-Mn2O3[55]
α-Mn2O3Cubic;
octahedral;
truncated octahedral
[Catalyst] = 4 g/L;
[Glycerol] = 20 g/L;
High catalytic activity (0.87 mmol/(h m2)) and high selectivity for glycerol (52.6%) was achieved by α-Mn2O3-truncated octahedralCo-exposed (001) and (111) facets of α-Mn2O3-truncated octahedral[60]
α-Mn2O3Octahedral;
truncated octahedral
180 mg of catalysts;
500 ppm of NO;
500 ppm of NH3;
5% v/v of O2;
N2 as balance gas;
36,000 h−1 of GSHV;
High NO turnover frequency ((3.6 ± 0.1) × 10−3 s−1) was achieved by α-Mn2O3-truncated octahedral
at 513 K
The exposure of a small fraction of (001) facets in α-Mn2O3-truncated octahedral[61]
α-Mn2O3Cubic;
octahedral;
truncated octahedral
100 mg of catalysts;
10 mg of soot;
5% v/v of O2;
0.25% v/v of NO;
N2 as balance gas;
9990 h−1 of GSHV;
96.3, 89.7, and 85.2% of soot combustion efficiencies were observed with the catalysis of α-Mn2O3-cubic, -truncated octahedral, -octahedralThe exposed (001) facet of cubic Mn2O3[56]
γ-MnOOHNanowires;
multi-branches;
nanorods
[Catalyst] = 0.3 g/L;
[PMS] = 12 mM;
[2,4-DCP] 2 = 100 mg/L;
pH = 7;
98%, 88%, and 55% removal of 2,4-DCP was achieved in γ-MnOOH nanowires, multi-branches, and nanorods activated PMS systems, separatelyHigher zeta-potential value of nanowires γ-MnOOH[59]
Mn3O4Nano-cubic;
nano-plate;
nano-octahedral
[Catalyst] = 0.2 g/L;
[PMS] = 0.65 mM;
[CIP] 3 = 10 mg/L;
pH = 7.7;
100% CIP removal in 80 min by Mn3O4 nano-octahedralLager surface Mn(IV) contents of Mn3O4 nano-octahedral[62]
1 Ref.: Reference; 2 2,4-DCP: 2,4-dichlorophenol; 3 CIP: ciprofloxacin.

3. Mechanisms of PMS/PDS Activation by Mn(III) (Oxyhydr)Oxides

3.1. Activation of PMS by Mn(III) (Oxyhydr)Oxides

The Mn(III) (oxyhydr)oxides/PMS system has been applied for the removal of a number of contaminants, such as phenol, bisphenol A, 2,4-dichlorophenol, ciprofloxacin, and organic dyes [62,63,64,65,66,67]. Different studies involving PMS activation by Mn(III) (oxyhydr)oxides are gathered in Table 3. According to the literature, the efficient degradation of organic pollutants is generally attributed to the generation of active species, such as SO 4 , HO , 1O2. The activation mechanisms of PMS by Mn(III) (oxyhydr)oxides are proposed, as shown in Figure 2. The simultaneous formation of Mn(II) and Mn(IV) and the conversion of Mn ions with different oxidation states explained well the good performance of Mn(III) (oxyhydr)oxides on PMS activation (Equations (1)–(4)) [44]. Except for the above-mentioned processes, the direct generation of HO   by Mn(III) activation of PMS was also reported by some researchers (Equation (5)) [62,64,66,68,69,70]. In comparison with SO 4 radical, the SO 5 radical has been regarded as a low oxidative activity for organic pollutants removal due to its low reduction potential (E0 = 1.10 V vs. NHE) [71]. Nevertheless, the transformation from SO 5 to SO 4 in Mn(III) (oxyhydr)oxides/PMS system still makes some contribution to the degradation of organic pollutants (Equation (6)) [72]. In addition, the conversion from SO 4 to HO in water should not be neglected (Equation (7)), especially, when the solution is in the alkaline environment (Equation (8)) [73].
Mn ( III ) + HSO 5   Mn ( IV ) + SO 4 + OH
Mn ( III ) + HSO 5   Mn ( II ) + SO 5 + H +
Mn ( II ) + HSO 5   Mn ( III ) + SO 4 + OH
Mn ( IV ) + HSO 5   Mn ( III ) + SO 5 + H +
Mn ( III ) + HSO 5   Mn ( IV ) + SO 4 2 + HO
SO 5 + SO 5   O 2 + 2   SO 4
SO 4   + H 2 O SO 4 2 + HO + H +
SO 4 + OH SO 4 2 + HO
In addition to the active radicals, the generation of non-radical species (such as 1O2) in the Mn(III) (oxyhydr)oxides-activated PMS system was also reported. For example, He et al. demonstrated the contribution of 1 O 2 for the degradation of 2,4-dichlorophenol in the γ-MnOOH/PMS system. The generation of 1 O 2 was attributed to two pathways including the decomposition of PMS and the reaction of O 2 with HO (Equations (9) and (10)) [59,74,75]. Chen et al. synthesized one new Mn3O4 nanodots-g-C3N4 nanosheet (Mn3O4/CNNS) and investigated its performance on PMS activation for 4-chlorophenol (4-CP) degradation [76]. The chemical scavenging tests and electron spin resonance (ESR) experiments confirmed the contribution of 1 O 2 for the removal of 4-CP. Furthermore, new pathways for the formation of 1 O 2 were reported in the Mn3O4/CNNS/PMS system. As shown in Equations (11)–(16), the reaction between SO 5 and H 2 O and the combination of O 2   with H 2 O can contribute to the formation of 1 O 2 [76].
HSO 5 + SO 5 2   HSO 4 + SO 4 2 +   1 O 2
O 2 + HO     1 O 2 + OH
2   SO 5 + H 2 O   2   HSO 4 + 1.5   1 O 2
HSO 5 + H 2 O   HSO 4 + H 2 O 2
H 2 O 2 H + + HO 2       pKa = 11.6
H 2 O 2 + HO H 2 O + HO 2
HO 2 H + + O 2       pKa = 4.88
2   O 2 + 2   H 2 O   H 2 O 2 +   1 O 2 + 2   OH
Currently, the Mn-based oxide composites have attracted increasing attention due to their various advantages, such as more oxygen vacancies, higher surface oxygen mobility, and enforced synergistic effects. For instance, Chen et al. prepared the Fe2O3/Mn2O3 composite and studied its activity on PMS activation for tartrazine (TTZ) degradation. The results showed that 97.3% removal of TTZ was achieved in 30 min in the Fe2O3/Mn2O3/PMS system. The efficient degradation of TTZ originated from the generation of active species (e.g., SO 4 , HO ) and the synergistic effect between iron and manganese ions [77]. The γ-MnOOH-coated nylon membrane was synthesized and applied in the activation of PMS towards the removal of 2,4-dichlorophenol (2,4-DCP). The deep removal of 2,4-DCP was explained by the synergetic “trap-and-zap” process, which improved the stability and catalytic reactivity of γ-MnOOH [63]. In conclusion, the activation of PMS by Mn(III) (oxyhydr)oxides, including pure Mn(III) oxides and Mn(III) containing composites, is favorable. The degradation of various pollutants in the Mn(III) (oxyhydr)oxides/PMS system can be achieved through the generation of active radicals and non-radical species.
Table 3. Summary of PMS activation by Mn(III) (oxyhydr)oxides.
Table 3. Summary of PMS activation by Mn(III) (oxyhydr)oxides.
CatalystsPollutantInitial ConditionsReactivityActive SpeciesRef.
Mn2O3Phenol[Catalyst] = 0.4 g/L;
[PMS] = 2 g/L;
[Phenol] = 25 mg/L;
100% removal of phenol in 60 min SO 4 [44]
Mn3O4Phenol[Catalyst] = 0.4 g/L;
[PMS] = 2 g/L;
[Phenol] = 25 mg/L;
100% removal of phenol in 20 min SO 4 [78]
Mn3O4 nanoparticleMethylene blue
(MB)
[Catalyst] = 0.12 g/L;
[PMS] = 0.94 g/L;
[MB] = 62 mg/L;
pH = 4;
86.71% removal of MB in 20 min SO 4 [64]
Mn3O4 nano-octahedralCiprofloxacin
(CIP)
[Catalyst] = 0.2 g/L;
[PMS] = 0.65 mM;
[CIP] = 10 mg/L;
pH = 7.7;
100% removal of CIP in 80 min SO 4
HO
[62]
yolk-shell Mn3O4Bisphenol A
(BPA)
[Catalyst] = 0.1 g/L;
[PMS] = 0.3 g/L;
[BPA] = 10 mg/L;
pH = 5.3;
87.7% of removal of BPA in 60 min SO 4
HO
[67]
3D hierarchical Mn3O4Phenol[Catalyst] = 0.2 g/L;
[PMS] = 6.5 mM;
[Phenol] = 20 ppm;
pH = 6.8;
100% removal of phenol in 60 min SO 4
HO
[66]
dumbbell-like Mn2O3Rhodamine B
(RhB)
[Catalyst] = 0.25 g/L;
[PMS] = 0.75 g/L;
[RhB] = 10 mg/L;
100% of removal of RhB in 30 min SO 4
HO
O 2
1 O 2
[65]
α-Mn2O3-cubicPhenol[Catalyst] = 0.4 g/L;
[PMS] = 2 g/L;
[Phenol] = 25 ppm;
100% removal of phenol in 1 h SO 4 [55]
γ-MnOOH nanowire2,4-dichlorophenol
(2,4-DCP)
[Catalyst] = 0.3 g/L;
[PMS] = 12 mM;
[2,4-DCP] = 100 mg/L;
pH = 7;
98% removal of 2,4-DCP in 6 h SO 4
HO
O 2
1 O 2
[59]
MnOOH@nylon2,4-DCP[Catalyst] = 0.76; mg/cm2;
[PMS] = 138 mg/L;
[2,4-DCP] = 25 mg/L;
pH = 6.0-6.4;
97.9% removal of 2,4-DCP in 2 h SO 4
HO
O 2
1 O 2
[63]
γ-MnOOH-rGOBentazone[Catalyst] = 0.075 g/L;
[PMS] = 0.615 g/L;
[Bentazone] = 10 mg/L;
pH = 7; sunlight;
96.1% removal of Bentazone in 90 min HO
1 O 2
[79]
Ce-Mn2O32,4-DCP[Catalyst] = 0.2 g/L;
[PMS] = 1.0 g/L;
[2,4-DCP] = 50 mg/L;
pH = 7;
100% removal of 2,4-DCP in 90 min SO 4
HO
1 O 2
[80]
Mn3O4-GOOrange II[Catalyst] = 50 mg/L;
[PMS] = 1.5 g/L;
[Orange II] = 30 mg/L;
pH = 7.0;
100% removal of Orange II in 120 min SO 4 [81]
Fe2O3/Mn2O3Tartrazine
(TTZ)
[Catalyst] = 0.6 g/L;
[PMS] = 0.8 g/L;
[TTZ] = 10 mg/L;
pH = 6.89;
97.3% removal of TTZ in 30 min SO 4
HO
[77]
Mn2O3@Mn5O84-chlorophenol
(4-CP)
[Catalyst] = 0.3 g/L;
[PMS] = 1.5 mM;
[4-CP] = 80 ppm;
100% removal of 4-CP in 60 min SO 4
HO
O 2
1 O 2
[82]
Mn3O4-MnO2CIP[Catalyst] = 0.1 g/L;
[PMS] = 1 mM;
[CIP] = 50 μM;
pH = 7.0 ± 0.1;
97.6% removal of CIP in 25 min SO 4
HO
[68]
Mn3O4/MOFRhB[Catalyst] = 0.4 g/L;
[PMS] = 0.3 g/L;
[RhB] = 10 mg/L;
pH = 5.18;
98% removal of RhB in 60 min SO 4
HO
[69]
Fe3O4/Mn3O4/GOMB[Catalyst] = 100 mg/L;
[PMS] = 0.3 g/L;
[MB] = 50 mg/L;
pH = 7;
98.8% removal of MB in 30 min SO 4

HO
[83]
Mn3O4/CNNS-1504-CP[Catalyst] = 0.3 g/L;
[PMS] = 1 mM;
[4-CP] = 50 mg/L;
pH = 6.89;
100% removal of 4-CP in 60 min1 O 2 [76]
α-Mn2O3@α-MnO2-350Phenol[Catalyst] = 0.4 g/L;
[PMS] = 2.0 g/L;
[Phenol] = 25 mg/L;
pH = 3-3.5;
100% removal of phenol in 25 min SO 4
HO
[84]
α-Mn2O3@α-MnO2-500Phenol[Catalyst] = 0.15 g/L;
[PMS] = 1 mM;
[Phenol] = 25 ppm;
100% removal of phenol in 70 min SO 4
HO
1 O 2
[85]
CuS/Fe2O3/Mn2O3CIP[Catalyst] = 0.6 g/L;
[PMS] = 0.6 g/L;
[CIP] = 10 mg/L;
pH = 5.84;
88% removal of CIP in 120 min SO 4
HO
[86]

3.2. Activation of PDS by Mn(III) (Oxyhydr)Oxides

Single or combined Mn(III) (oxyhydr)oxides have been employed to activate PDS to remove different organic pollutants, such as phenol, p-chloroaniline (PCA), 2,4-dichlorophenol (2,4-DCP), and organic dyes (Table 4). The activation pathway of PDS varies with the different types of Mn(III) (oxyhydr)oxides (Figure 3). For example, Shabanloo et al. reported the generation of active SO 4   radicals in the nano-Mn3O4/PDS system [87]. Since both Mn(II) and Mn(III) species are identified in the Mn3O4 structure, the formation of SO 4 was mainly attributed to the activation of PDS by Mn(II) (Equation (17)). In contrast, the persulfate radical ( S 2 O 8 ) was produced by the reaction of PDS and Mn(III) (Equation (18)). For the system of Mn2O3/PDS, it is believed that the singlet oxygen (1 O 2 ) was the primary active species that was responsible for the degradation of organic pollutants [88]. As demonstrated by Khan et al., one complex Mn(III/IV)- OS 2 O 7 was formed between PDS and Mn2O3 through the inner-sphere interaction. Then, another S 2 O 8 2 was decomposed by Mn(III/IV)- OS 2 O 7 to generate HO 2 / O 2   radicals. The 1 O 2 was finally formed from the direct oxidation of O 2   by Mn(IV)- OS 2 O 7 or the recombination of HO 2 and O 2 (Equations (19)–(20)). The pathway of 1 O 2 formation in the system of A-Mn2O3/PDS is comparable to the approach of producing 1 O 2 in the β-MnO2/PDS system in which the important metastable manganese intermediate was first formed through the complex reaction between the hydroxyl group (-OH) and cleavaged S 2 O 8 2 [41]. Therefore, the hydroxyl group on the surface of manganese oxides plays a significant role in PDS activation.
Mn ( II ) + S 2 O 8 2 Mn ( III ) + SO 4 + SO 4 2 .
Mn ( III ) + S 2 O 8 2 Mn ( II ) + S 2 O 8
Mn ( IV ) OS 2 O 7 + O 2 + OH Mn ( III ) OH + 2   SO 4 2 +   1 O 2
O 2 + HO 2     1 O 2 + H O 2
In comparison with Mn3O4 and Mn2O3, γ-MnOOH presents more -OH groups on the surface, leading to the high efficiency in PDS activation. For instance, Li et al. reported that γ-MnOOH exhibited higher reactivity in PDS activation for phenol oxidation in comparison with Mn2O3 and Mn3O4 [89]. The authors reported that the degradation efficiency of phenol in the γ-MnOOH/PDS system was pH-dependent. Under the basic condition (pH 11), phenol was efficiently removed due to the generation of SO 4 and HO radicals. However, at pH 3 and 7, the oxidative intermediate ( Mn ( III ) O 3 SOOSO 3 ) was believed to be responsible for the removal of phenol. Although the mentioned report explained well the oxidation performance of γ-MnOOH/PMS for phenol removal, the information regarding the mechanism of PDS activation on the surface of γ-MnOOH was not given in detail. Considering this, Xu et al. conducted a further investigation focusing on the catalytic mechanism of PDS by γ-MnOOH [90]. Based on the results of chemical scavenging and ESR experiments, a non-radical mechanism was proposed. Generally, the non-radical mechanism in PS activation was attributed to three aspects—the generation of 1 O 2 , the electron transfer process, and the catalyst surface-activated intermediates [91,92,93,94,95]. However, the 1 O 2   production and electron transfer process mechanism were excluded according to the results of ESR and linear sweep voltammetry (LSV) experiments. Therefore, the γ-MnOOH surface-activated PDS molecules were verified as the main active species for the degradation of PCA. Figure 4 shows the formation of active PDS molecules on the surface of γ-MnOOH.
The activation of PDS by Mn(III) (oxyhydr)oxide composites for pollutant degradation was also reported [96,97,98]. For instance, Liu et al. synthesized the carbon-coated Mn3O4 composite (Mn3O4/C) and investigated the reactivity in the presence of PDS for 2,4-dichlorophenol (2,4-DCP) degradation [96]. The results showed that 95% of 2,4-DCP removal was reached in 140 min and the enhanced degradation was attributed to the existence of the defective edges of the carbon layer, which facilitated the attraction and activation of PDS. Rizal et al. prepared Ag/Mn3O4 and Ag/Mn3O4/graphene composites and studied the degradation efficiency of methylene blue (MB) by the synthesized catalysts activated PDS in the presence of visible light [97]. The results showed that 40 mg/L of MB was completely removed in 30 min by the system of Ag/Mn3O4/graphene + PDS under visible light. The enhanced degradation of MB was attributed to the hampered electron-hole recombination due to the loading of Ag and graphene. Furthermore, the studies regarding the application of modified Mn2O3 in oxidants (such as PMS, H 2 O 2 ) activation for contaminants removal were also reported [84,99,100,101]. For example, Saputra et al. prepared an egg-shaped core/shell α-Mn2O3@α-MnO2 catalyst via a hydrothermal process and investigated the catalytic activity of α-Mn2O3@α-MnO2 in heterogeneous Oxone® activation for phenol degradation [84]. The loaded α-MnO2 improved the generation of Mn(III) species through the reaction with PMS. The amount of SO 4 and HO was then increased leading to the enhanced degradation of phenol. The efficient degradation of organic dye pollutants (such as Rhodamine B (RhB) and Congo Red (CR)) by bimetallic Mn2O3-Co3O4/carbon catalyst activated Fenton-like reaction was also reported [100]. The superior reactivity of Mn2O3-Co3O4/C catalyst in H 2 O 2 activation for RB and CR degradation was attributed to the good synergistic effect between Co3O4 and Mn2O3 as well as the interaction between metal oxides and carbon. However, the investigation regarding the activation of PDS by modified α-Mn2O3 has been less reported. The same effect was also observed for the γ-MnOOH-based composites. This might be attributed to the distinct activation way of PDS by α-Mn2O3 or γ-MnOOH compared with Mn3O4.
In summary, Mn3O4 can activate PDS to generate SO 4 through radical mechanisms, while the activation of PDS by α-Mn2O3 and γ-MnOOH is processed in a non-radical mechanism with the generation of 1O2 and catalyst surface-activated PDS substances. For the activation of PDS by Mn(III) (oxyhydr)oxides composites, the Mn3O4-based composites have shown good catalytic performance in PDS activation for pollutant degradation. In comparison, the activation of PDS by modified α-Mn2O3 or γ-MnOOH catalysts needs to be further investigated.
Table 4. Summary of PDS activation by Mn(III) (oxyhydr)oxides.
Table 4. Summary of PDS activation by Mn(III) (oxyhydr)oxides.
CatalystsPollutantInitial ConditionsReactivityActive SpeciesRef.
γ-MnOOHP-chloroaniline
(PCA)
[Catalyst] = 0.4 g/L;
[PDS] = 2.5 mM;
[PCA] = 0.5 mM;
pH = 4.2;
100% removal of PCA in 180 minγ-MnOOH-PDS complex[90]
A-Mn2O3Phenol[Catalyst] = 0.2 g/L;
[PDS] = 2 mM;
[Phenol] = 12 ppm;
pH = 3.2;
100% removal of phenol in 70 min1 O 2 [88]
Mn3O4 nanoparticleAcid Blue 113
(AB113)
[Catalyst] = 57.69 mg/L;
[PDS] = 61.46 mg/L;
[AB113] = 50 mg/L;
pH = 3;
96.7% removal of AB113 in 60 min SO 4
HO
[102]
γ-MnOOHPhenol[Catalyst] = 1 g/L;
[PDS] = 2 g/L;
[Phenol] = 100 mg/L;
pH = 7;
91.86% removal of phenol in 360 minγ-MnOOH-PDS complex[89]
Nano-Mn3O4Furfural[Catalyst] = 1.2 g/L;
[PDS] = 6.34 mM;
[Furfural] = 50 mg/L;
pH = 4.82;
91.14% of furfural removal in 60 min SO 4 [87]
Ag/Mn3O4-5 GMB[Catalyst] = 0.5 g/L;
[PDS] = 12 mM;
[MB] = 40 mg/L;
pH = 3;
visible-light;
100% of MB removal in 30 min SO 4
HO
[97]
Mn2O3/Mn3O4/MnO2-10Orange II[Catalyst] = 0.4 g/L;
[PDS] = 2 g/L;
[Orange II] = 20 mg/L;
95% removal of Orange II in 50 min SO 4
HO
[103]
0.5-Mn3O4/C-T42,4-DCP[Catalyst] = 0.2 g/L;
[PDS] = 2 g/L;
[2,4-DCP] = 100 mg/L;
pH = 6.37;
95% removal of 2,4-DCP in 140 min SO 4
HO
1 O 2
[96]
γ-Fe2O3/Mn3O4RhB[Catalyst] = 50 mg/L;
[PDS] = 50 mg/L;
[RhB] = 10 mg/L;
pH = 4.5;
97.5% removal of RhB in 150 min SO 4
HO
[98]

4. Influence Factors for Mn(III) (Oxyhydr)Oxides Reactivity

4.1. The Effect of pH

The Mn(III) (oxyhydr)oxides-mediated activation of PDS/PMS can be affected by solution pH in different ways. For example, influencing the property of charge on the surface of the catalysts, changing the ionic forms of PDS/PMS and pollutant molecules, as well as altering the reduction potential of active radicals.
First, the solution pH can affect the interaction between catalyst and PDS/PMS and pollutants through changing the electrostatic effect. The point of zero charges (PZC) of the catalyst and the acid dissociation constant (pKa) of radical precursors and contaminants are two important parameters that are used to recognize the charge type on the surface of the catalysts and the ionic situation of oxidants and pollutants in solution. For instance, when the solution pH is equal to the PZC value of the catalyst, the amounts of positive and negative charges on the surface of the catalyst are equal (i.e., the surface charge of the catalyst is zero). When the solution pH is higher than the PZC value of the catalyst, the surface charges of the catalyst are negative. On the contrary, if the solution pH is lower than the catalyst PZC value, the surface of the catalyst will be positively charged [104]. The same situation is suitable for the analysis of the ionic form of oxidants and pollutants. The PZC values of commonly used Mn(III) (oxyhydr)oxides and the pKa values of PMS/PDS, and some typical pollutants, are summarized in Table 5. The impacts of solution pH on the interaction between Mn(III) (oxyhydr)oxides and PDS/PMS and pollutants have been reported. For example, Zhao et al. reported that the adsorption and degradation of ciprofloxacin (CIP) by the synthesized Mn3O4-MnO2 composite were facilitated at neutral pH solution [68]. The results were explained by the enhanced electrostatic attraction between Mn3O4-MnO2 and CIP. The PZC value of the Mn3O4-MnO2 composite was measured at 2.5; thus, in the solution pH 7, the surface of the catalyst was negatively charged. In comparison, the pKa of CIP was 8.7–10.58, leading to the formation of positively charged CIP ions in the neutral pH solution. Therefore, the electrostatic attraction between the negative catalyst and the positive CIP improved, resulting in a facilitating degradation of CIP. The same phenomenon was also reported in the studies of PDS activation by γ-MnOOH/α-Mn2O3 for pollutant degradation [88,90].
Second, the transformation of radicals also influenced the reactivity of Mn(III) (oxyhydr)oxides for pollutant degradation. For instance, the reported conversion of SO 4 to HO under the basic solution (as shown in (Equation (8)) can have a significant impact. Since the reduction potential value of HO under natural pH is lower than that in acidic solution (1.8 vs. 2.7V) [105], and the lifetime of HO is shorter than SO 4 (20 ns vs. 30–40 μs) [106]; thus, the transformation from SO 4 (E = 2.6 V) to HO under alkaline solution might lead to a decrease of pollutant degradation. In addition, the leaching of Mn2+ from Mn(III) (oxyhydr)oxides in an acidic condition also should be taken into consideration for the activation of sulfate compounds (PMS/PDS).
Table 5. The PZC values of Mn(III) (oxyhydr)oxides and pKa values of PMS/PDS and pollutants.
Table 5. The PZC values of Mn(III) (oxyhydr)oxides and pKa values of PMS/PDS and pollutants.
CatalystsPZCReference
α-Mn2O34.7[88,107]
γ-MnOOH3.4[90]
Mn3O45.6–7.34[68,87,102]
OxidantspKaReference
PMS9.4[108]
PDS−3.5[109]
PollutantspKaReference
Phenol9.98[110]
Bisphenol A9.6–10.2[111]
2,4-dichlorophenol9.4[82]
Ciprofloxacin8.70–10.58[68,112]
p-Chloroaniline4.2[90,113]
4-Chlorophenol9.29[114]
Orange II11.4[103]

4.2. The Effect of Inorganic Anions

Inorganic anions are ubiquitous in various aquatic compartments. It is reported that inorganic anions can suppress the degradation of pollutants in Mn(III) (oxyhydr) oxides-activated PMS/PDS systems through competing with pollutants for radicals. Thus, to evaluate the applicability of the Mn(III) (oxyhydr)oxides + PMS/PDS system in different water matrices, the influence of inorganic anions on the removal of pollutants has been investigated by many researchers [63,79,86,88,97]. In this section, the effect of inorganic anions, such as carbonate/bicarbonate ions ( CO 3 2 / HCO 3 ), chloride ions ( Cl ), and nitrate ( NO 3 )/nitrite ions ( NO 2 ) on the reactivity of Mn(III) (oxyhydr)oxides was summarized.
Carbonate ( CO 3 2 )/bicarbonate ( HCO 3 ) can react with SO 4 and HO to generate less reactive carbonate radical ( CO 3 ) and bicarbonate radical ( HCO 3 ) (Equations (21)–(25)) leading to the inhibited degradation of pollutants [115]. However, although the redox potential of CO 3 is low (1.59 V vs. NHE), it can still selectively degrade some organic pollutants with a reaction rate of 103–109 M−1s−1 [116,117]. In addition, the presence of carbonate and bicarbonate ions can affect the stability of oxidants. For example, PDS can be activated by HCO 3 to generated percarbonate ( HCO 4 ) (Equation (26)) [118]. Similarly, PMS can be catalyzed by both CO 3 2 and HCO 3 to form active radicals and HCO 4 (Equations (27)–(29)). Furthermore, the solution pH can be changed in the presence of carbonate/bicarbonate ions, which can affect the reactivity of Mn(III) (oxyhydr)oxides in PMS/PDS activation as discussed in Section 4.1.
SO 4 + CO 3 2     SO 4 2 + CO 3
SO 4 + HCO 3     SO 4 2 + HCO 3
HO + CO 3 2     OH + CO 3
HO + HCO 3   H 2 O + HCO 3
HCO 3 = H + + CO 3   pKa = 9.6
S 2 O 8 2 + HCO 3 + 2   OH     HCO 4 + 2   SO 4 2 + H 2 O
HSO 5 + CO 3 2 + H +     SO 4 + 2   OH + CO 2
HSO 5 + HCO 3     SO 4 + 2   OH   + CO 2
HSO 5 + HCO 3   SO 4 2 + HCO 4 + H +
Chloride ion ( Cl ) exists widely in various water bodies including surface water, groundwater, and industrial wastewater [119]. The influence of Cl on the degradation of organic pollutants by sulfate radical-based AOPs (SR-AOP) was reported by previous studies [120,121,122,123]. Generally, Cl can react with SO 4 to generate Cl , which can react with another Cl to form Cl 2 (Equations (30)–(31)) [122]. Both Cl and Cl 2 have low reduction potentials (E0 = 2.4 and 2.0 V) in comparison with SO 4 , thus the consumption of SO 4 by Cl leads to the decrease of organic pollutant degradation [124,125]. However, Cl was believed to own higher selectivity on electron-rich compounds than nonselective SO 4 , which can offset the negative effect of Cl on SO 4 [126]. Therefore, the conflicting effect of Cl   on organic pollutants in SR-AOP might be attributed to the different reactivity of pollutants with Cl and Cl 2 . In addition, the reactivity of HO   can also be suppressed by Cl due to the formation of low active radical ClOH (Equation (32)) [127].
SO 4 + Cl   SO 4 2 + Cl
Cl   + Cl   Cl 2
HO + Cl   ClOH
Nitrate ( NO 3 ) and nitrite ( NO 2 ) can be commonly found in various water matrices [119]. Both NO 3 and NO 2 are able to react with SO 4 to generate low reactive NO 3 (E0 = 2.3–2.5 V) and NO 2 radicals (E0 = 1.03 V) (Equations (33)–(34)) [25]. The reaction rate of SO 4 with NO 3 and NO 2 are 5 × 104 M−1s−1 and 8.8 × 108 M−1s−1, separately [45]. Thus, NO 2 , compared with NO 3 , has higher reactivity in SO 4 consumption. In addition and in a similar way, NO 2 was also reported as the sink of HO radicals (Equation (35)) [128].
SO 4 + NO 3   SO 4 2 + NO 3
SO 4 + NO 2   SO 4 2 + NO 2
HO + NO 2   OH + NO 2

5. Summary and Outlooks

This review summarized the activation of PMS and PDS by manganese(III) (oxyhydr)oxides for the degradation of recalcitrant pollutants. The desirable morphologies and facets (e.g., cubic structure with (001) facet exposure) can effectively enhance the reactivity of Mn(III) (oxyhydr)oxides in the activation of PDS and PMS. Mn(III) (oxyhydr)oxides showed different reactivity in radical precursors activation. Specifically, both radical (for example, sulfate and hydroxyl radical) and non-radical (such as singlet oxygen) were generated in the Mn(III) (oxyhydr)oxide-activated PMS system. The activation of PDS by α-Mn2O3 and γ-MnOOH were mainly through the formation of singlet oxygen and the catalyst surface activated complex. The activity of Mn(III) (oxyhydr)oxides in PDS and PMS activation can be influenced by the solution pH due to the occurrence of the electrostatic effect. Moreover, the inhibition effect of inorganic anions (such as carbonate/bicarbonate ions, chloride ions, and nitrate/nitrite ions) on the catalytic performance of Mn(III) (oxyhydr)oxides were discussed in detail.
Given this comprehensive summary, some future outlooks are proposed.
Although previous studies already identified the generation of 1O2 in α-Mn2O3/PDS system using the ESR and quenching experiments, the detailed catalytic process of PDS on the surface of Mn2O3 remains elusive. Further studies are needed for a better understanding of the activation mechanism of PDS by α-Mn2O3. Second, detailed studies are required to exploit the potential application of α-Mn2O3 or γ-MnOOH-based composites in PDS activation to understand the synergistic performance of α-Mn2O3 or γ-MnOOH with other loaded materials (such as active carbon, graphene, and transition metals). This will open up new research avenues in the field of water remediation technologies, with the aim to improve the reactivity of α-Mn2O3/γ-MnOOH in PDS activation.
The high-scale or industrial application of SR-AOP seems difficult to implement, and that merits being resolved. The development of new modeling approaches that account for the upscaling of different involved reactions and the complexity of heterogeneous reactions at Mn-oxides/water interfaces becomes urgent. More experimental work is also needed to develop new Mn-bearing oxides supported with high catalytic efficiency, suitable for industrial applications, and yet are relevant from both economic and environmental points of view.

Author Contributions

Draft preparation, D.J.; conceptualization, D.J., M.B.; revised the paper D.J., G.M., M.B., K.H. All authors have read and agreed to the published version of the manuscript.

Funding

No funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

We gratefully acknowledge the Chinese Scholarship Council of China for providing financial support for Daqing Jia. We acknowledge the program PAI (Pack Ambition Recherche) SOLDE from the Region Auvergne Rhône Alpes for the financial support of D.J. in this project.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Rathi, B.S.; Kumar, P.S.; Show, P.-L. A Review on Effective Removal of Emerging Contaminants from Aquatic Systems: Current Trends and Scope for Further Research. J. Hazard. Mater. 2021, 409, 124413. [Google Scholar] [CrossRef]
  2. Kasonga, T.K.; Coetzee, M.A.A.; Kamika, I.; Ngole-Jeme, V.M.; Benteke Momba, M.N. Endocrine-Disruptive Chemicals as Contaminants of Emerging Concern in Wastewater and Surface Water: A Review. J. Environ. Manag. 2021, 277, 111485. [Google Scholar] [CrossRef] [PubMed]
  3. Chen, L.; Fu, W.; Tan, Y.; Zhang, X. Emerging Organic Contaminants and Odorous Compounds in Secondary Effluent Wastewater: Identification and Advanced Treatment. J. Hazard. Mater. 2021, 408, 124817. [Google Scholar] [CrossRef] [PubMed]
  4. Sui, Q.; Jiang, C.; Zhang, J.; Yu, D.; Chen, M.; Wang, Y.; Wei, Y. Does the Biological Treatment or Membrane Separation Reduce the Antibiotic Resistance Genes from Swine Wastewater through a Sequencing-Batch Membrane Bioreactor Treatment Process. Environ. Int. 2018, 118, 274–281. [Google Scholar] [CrossRef]
  5. Verma, S.; Daverey, A.; Sharma, A. Slow Sand Filtration for Water and Wastewater Treatment—A Review. Environ. Technol. Rev. 2017, 6, 47–58. [Google Scholar] [CrossRef]
  6. Rebosura, M.; Salehin, S.; Pikaar, I.; Keller, J.; Sharma, K.; Yuan, Z. The Impact of Primary Sedimentation on the Use of Iron-Rich Drinking Water Sludge on the Urban Wastewater System. J. Hazard. Mater. 2021, 402, 124051. [Google Scholar] [CrossRef] [PubMed]
  7. Huang, H.; Chen, Y.; Jiang, Y.; Ding, L. Treatment of Swine Wastewater Combined with MgO-Saponification Wastewater by Struvite Precipitation Technology. Chem. Eng. J. 2014, 254, 418–425. [Google Scholar] [CrossRef]
  8. Lv, M.; Zhang, Z.; Zeng, J.; Liu, J.; Sun, M.; Yadav, R.S.; Feng, Y. Roles of Magnetic Particles in Magnetic Seeding Coagulation-Flocculation Process for Surface Water Treatment. Sep. Purif. Technol. 2019, 212, 337–343. [Google Scholar] [CrossRef]
  9. Teh, C.Y.; Budiman, P.M.; Shak, K.P.Y.; Wu, T.Y. Recent Advancement of Coagulation–Flocculation and Its Application in Wastewater Treatment. Ind. Eng. Chem. Res. 2016, 55, 4363–4389. [Google Scholar] [CrossRef]
  10. Wei, H.; Gao, B.; Ren, J.; Li, A.; Yang, H. Coagulation/Flocculation in Dewatering of Sludge: A Review. Water Res. 2018, 143, 608–631. [Google Scholar] [CrossRef]
  11. Béguin, P.; Aubert, J.-P. The Biological Degradation of Cellulose. FEMS Microbiol. Rev. 1994, 13, 25–58. [Google Scholar] [CrossRef]
  12. Durai, G.; Rajasimman, M. Biological Treatment of Tannery Wastewater—A Review. J. Environ. Sci. Technol. 2010, 4, 1–17. [Google Scholar] [CrossRef] [Green Version]
  13. Li, Y.; Dong, H.; Li, L.; Tang, L.; Tian, R.; Li, R.; Chen, J.; Xie, Q.; Jin, Z.; Xiao, J.; et al. Recent Advances in Waste Water Treatment through Transition Metal Sulfides-Based Advanced Oxidation Processes. Water Res. 2021, 192, 116850. [Google Scholar] [CrossRef] [PubMed]
  14. Hu, X.; Wang, X.; Ban, Y.; Ren, B. A Comparative Study of UV–Fenton, UV–H2O2 and Fenton Reaction Treatment of Landfill Leachate. Environ. Technol. 2011, 32, 945–951. [Google Scholar] [CrossRef] [PubMed]
  15. Song, S.; Wang, Y.; Shen, H.; Zhang, J.; Mo, H.; Xie, J.; Zhou, N.; Shen, J. Ultrasmall Graphene Oxide Modified with Fe3O4 Nanoparticles as a Fenton-Like Agent for Methylene Blue Degradation. ACS Appl. Nano Mater. 2019, 2, 7074–7084. [Google Scholar] [CrossRef]
  16. Lan, H.; Wang, F.; Lan, M.; An, X.; Liu, H.; Qu, J. Hydrogen-Bond-Mediated Self-Assembly of Carbon-Nitride-Based Photo-Fenton-like Membranes for Wastewater Treatment. Environ. Sci. Technol. 2019, 53, 6981–6988. [Google Scholar] [CrossRef]
  17. Vermilyea, A.W.; Voelker, B.M. Photo-Fenton Reaction at Near Neutral PH. Environ. Sci. Technol. 2009, 43, 6927–6933. [Google Scholar] [CrossRef]
  18. Ding, J.; Sun, Y.-G.; Ma, Y.-L. Highly Stable Mn-Doped Metal-Organic Framework Fenton-Like Catalyst for the Removal of Wastewater Organic Pollutants at All Light Levels. ACS Omega 2021, 6, 2949–2955. [Google Scholar] [CrossRef]
  19. Jaafarzadeh, N.; Takdastan, A.; Jorfi, S.; Ghanbari, F.; Ahmadi, M.; Barzegar, G. The Performance Study on Ultrasonic/Fe3O4/H2O2 for Degradation of Azo Dye and Real Textile Wastewater Treatment. J. Mol. Liq. 2018, 256, 462–470. [Google Scholar] [CrossRef]
  20. Mahamuni, N.N.; Adewuyi, Y.G. Advanced Oxidation Processes (AOPs) Involving Ultrasound for Waste Water Treatment: A Review with Emphasis on Cost Estimation. Ultrason. Sonochem. 2010, 17, 990–1003. [Google Scholar] [CrossRef] [PubMed]
  21. Nidheesh, P.V.; Gandhimathi, R. Trends in Electro-Fenton Process for Water and Wastewater Treatment: An Overview. Desalination 2012, 299, 1–15. [Google Scholar] [CrossRef]
  22. Melin, V.; Salgado, P.; Thiam, A.; Henríquez, A.; Mansilla, H.D.; Yáñez, J.; Salazar, C. Study of Degradation of Amitriptyline Antidepressant by Different Electrochemical Advanced Oxidation Processes. Chemosphere 2021, 274, 129683. [Google Scholar] [CrossRef]
  23. Sgroi, M.; Anumol, T.; Vagliasindi, F.G.A.; Snyder, S.A.; Roccaro, P. Comparison of the New Cl2/O3/UV Process with Different Ozone- and UV-Based AOPs for Wastewater Treatment at Pilot Scale: Removal of Pharmaceuticals and Changes in Fluorescing Organic Matter. Sci. Total Environ. 2021, 765, 142720. [Google Scholar] [CrossRef] [PubMed]
  24. Chen, H.; Wang, J. Degradation and Mineralization of Ofloxacin by Ozonation and Peroxone (O3/H2O2) Process. Chemosphere 2021, 269, 128775. [Google Scholar] [CrossRef] [PubMed]
  25. Giannakis, S.; Lin, K.-Y.A.; Ghanbari, F. A Review of the Recent Advances on the Treatment of Industrial Wastewaters by Sulfate Radical-Based Advanced Oxidation Processes (SR-AOPs). Chem. Eng. J. 2021, 406, 127083. [Google Scholar] [CrossRef]
  26. Duan, X.; Yang, S.; Wacławek, S.; Fang, G.; Xiao, R.; Dionysiou, D.D. Limitations and Prospects of Sulfate-Radical Based Advanced Oxidation Processes. J. Environ. Chem. Eng. 2020, 8, 103849. [Google Scholar] [CrossRef]
  27. Solís, R.R.; Rivas, F.J.; Chávez, A.M.; Dionysiou, D.D. Peroxymonosulfate/Solar Radiation Process for the Removal of Aqueous Microcontaminants. Kinetic Modeling, Influence of Variables and Matrix Constituents. J. Hazard. Mater. 2020, 400, 123118. [Google Scholar] [CrossRef] [PubMed]
  28. Chen, C.; Feng, H.; Deng, Y. Re-Evaluation of Sulfate Radical Based–Advanced Oxidation Processes (SR-AOPs) for Treatment of Raw Municipal Landfill Leachate. Water Res. 2019, 153, 100–107. [Google Scholar] [CrossRef] [PubMed]
  29. Anipsitakis, G.P.; Dionysiou, D.D. Transition Metal/UV-Based Advanced Oxidation Technologies for Water Decontamination. Appl. Catal. B Environ. 2004, 54, 155–163. [Google Scholar] [CrossRef]
  30. Ao, X.; Liu, W. Degradation of Sulfamethoxazole by Medium Pressure UV and Oxidants: Peroxymonosulfate, Persulfate, and Hydrogen Peroxide. Chem. Eng. J. 2017, 313, 629–637. [Google Scholar] [CrossRef]
  31. Lin, H.; Li, S.; Deng, B.; Tan, W.; Li, R.; Xu, Y.; Zhang, H. Degradation of Bisphenol A by Activating Peroxymonosulfate with Mn0.6Zn0.4Fe2O4 Fabricated from Spent Zn-Mn Alkaline Batteries. Chem. Eng. J. 2019, 364, 541–551. [Google Scholar] [CrossRef]
  32. Anipsitakis, G.P.; Dionysiou, D.D. Radical Generation by the Interaction of Transition Metals with Common Oxidants. Environ. Sci. Technol. 2004, 38, 3705–3712. [Google Scholar] [CrossRef]
  33. Huang, W.; Bianco, A.; Brigante, M.; Mailhot, G. UVA-UVB Activation of Hydrogen Peroxide and Persulfate for Advanced Oxidation Processes: Efficiency, Mechanism and Effect of Various Water Constituents. J. Hazard. Mater. 2018, 347, 279–287. [Google Scholar] [CrossRef] [PubMed]
  34. Gabet, A.; Métivier, H.; de Brauer, C.; Mailhot, G.; Brigante, M. Hydrogen Peroxide and Persulfate Activation Using UVA-UVB Radiation: Degradation of Estrogenic Compounds and Application in Sewage Treatment Plant Waters. J. Hazard. Mater. 2021, 405, 124693. [Google Scholar] [CrossRef] [PubMed]
  35. Ahn, Y.-Y.; Choi, J.; Kim, M.; Kim, M.S.; Lee, D.; Bang, W.H.; Yun, E.-T.; Lee, H.; Lee, J.-H.; Lee, C.; et al. Chloride-Mediated Enhancement in Heat-Induced Activation of Peroxymonosulfate: New Reaction Pathways for Oxidizing Radical Production. Environ. Sci. Technol. 2021, 55, 5382–5392. [Google Scholar] [CrossRef]
  36. Hu, J.; Zeng, X.; Yin, Y.; Liu, Y.; Li, Y.; Hu, X.; Zhang, L.; Zhang, X. Accelerated Alkaline Activation of Peroxydisulfate by Reduced Rubidium Tungstate Nanorods for Enhanced Degradation of Bisphenol A. Environ. Sci. Nano 2020, 7, 3547–3556. [Google Scholar] [CrossRef]
  37. Rodríguez-Chueca, J.; Giannakis, S.; Marjanovic, M.; Kohantorabi, M.; Gholami, M.R.; Grandjean, D.; de Alencastro, L.F.; Pulgarín, C. Solar-Assisted Bacterial Disinfection and Removal of Contaminants of Emerging Concern by Fe2+-Activated HSO5 vs. S2O82− in Drinking Water. Appl. Catal. B Environ. 2019, 248, 62–72. [Google Scholar] [CrossRef]
  38. Oh, W.-D.; Lim, T.-T. Design and Application of Heterogeneous Catalysts as Peroxydisulfate Activator for Organics Removal: An Overview. Chem. Eng. J. 2019, 358, 110–133. [Google Scholar] [CrossRef]
  39. Jia, D.; Li, Q.; Hanna, K.; Mailhot, G.; Brigante, M. Efficient Removal of Estrogenic Compounds in Water by MnIII-Activated Peroxymonosulfate: Mechanisms and Application in Sewage Treatment Plant Water. Environ. Pollut. 2021, 288, 117728. [Google Scholar] [CrossRef]
  40. Taujale, S.; Baratta, L.R.; Huang, J.; Zhang, H. Interactions in Ternary Mixtures of MnO2, Al2O3, and Natural Organic Matter (NOM) and the Impact on MnO2 Oxidative Reactivity. Environ. Sci. Technol. 2016, 50, 2345–2353. [Google Scholar] [CrossRef] [Green Version]
  41. 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] [PubMed]
  42. Zhou, Z.-G.; Du, H.-M.; Dai, Z.; Mu, Y.; Tong, L.-L.; Xing, Q.-J.; Liu, S.-S.; Ao, Z.; Zou, J.-P. Degradation of Organic Pollutants by Peroxymonosulfate Activated by MnO2 with Different Crystalline Structures: Catalytic Performances and Mechanisms. Chem. Eng. J. 2019, 374, 170–180. [Google Scholar] [CrossRef]
  43. Huang, J.; Dai, Y.; Singewald, K.; Liu, C.-C.; Saxena, S.; Zhang, H. Effects of MnO2 of Different Structures on Activation of Peroxymonosulfate for Bisphenol A Degradation under Acidic Conditions. Chem. Eng. J. 2019, 370, 906–915. [Google Scholar] [CrossRef]
  44. Saputra, E.; Muhammad, S.; Sun, H.; Ang, H.-M.; Tadé, M.O.; Wang, S. Manganese Oxides at Different Oxidation States for Heterogeneous Activation of Peroxymonosulfate for Phenol Degradation in Aqueous Solutions. Appl. Catal. B Environ. 2013, 142–143, 729–735. [Google Scholar] [CrossRef] [Green Version]
  45. Huang, J.; Zhang, H. Mn-Based Catalysts for Sulfate Radical-Based Advanced Oxidation Processes: A Review. Environ. Int. 2019, 133, 105141. [Google Scholar] [CrossRef]
  46. Liu, W.; Sutton, N.B.; Rijnaarts, H.H.M.; Langenhoff, A.A.M. Pharmaceutical Removal from Water with Iron- or Manganese-Based Technologies: A Review. Crit. Rev. Environ. Sci. Technol. 2016, 46, 1584–1621. [Google Scholar] [CrossRef] [Green Version]
  47. Remucal, C.K.; Ginder-Vogel, M. A Critical Review of the Reactivity of Manganese Oxides with Organic Contaminants. Environ. Sci. Proc. Impacts 2014, 16, 1247. [Google Scholar] [CrossRef] [PubMed]
  48. Wu, P.; Jin, X.; Qiu, Y.; Ye, D. Recent Progress of Thermocatalytic and Photo/Thermocatalytic Oxidation for VOCs Purification over Manganese-Based Oxide Catalysts. Environ. Sci. Technol. 2021, 55, 4268–4286. [Google Scholar] [CrossRef]
  49. Huang, J.; Zhong, S.; Dai, Y.; Liu, C.-C.; Zhang, H. Effect of MnO2 Phase Structure on the Oxidative Reactivity toward Bisphenol A Degradation. Environ. Sci. Technol. 2018, 52, 11309–11318. [Google Scholar] [CrossRef]
  50. Post, J.E. Manganese Oxide Minerals: Crystal Structures and Economic and Environmental Significance. Proc. Natl. Acad. Sci. USA 1999, 96, 3447–3454. [Google Scholar] [CrossRef] [Green Version]
  51. Ghosh, S.K. Diversity in the Family of Manganese Oxides at the Nanoscale: From Fundamentals to Applications. ACS Omega 2020, 5, 25493–25504. [Google Scholar] [CrossRef] [PubMed]
  52. Geller, S. Structure of α-Mn2O3, (Mn0.983Fe0.017)2O3 and (Mn0.37Fe0.63)2O3 and Relation to Magnetic Ordering. Acta Crystallogr. B. Struct. Sci. Cryst. Eng. Mater. 1971, 27, 821–828. [Google Scholar] [CrossRef]
  53. Baron, V.; Gutzmer, J.; Rundlof, H.; Tellgren, R. The Influence of Iron Substitution in the Magnetic Properties of Hausmannite, Mn2+ ( Fe , Mn ) 2 3 + O4. Am. Mineral. 1998, 83, 786–793. [Google Scholar] [CrossRef]
  54. Buerger, M.J. The Symmetry and Crystal Structure of Manganite, Mn (OH)O. Z. Kristallogr. Cryst. Mater. 1936, 95, 163–174. [Google Scholar] [CrossRef]
  55. Saputra, E.; Muhammad, S.; Sun, H.; Ang, H.-M.; Tadé, M.O.; Wang, S. Shape-Controlled Activation of Peroxymonosulfate by Single Crystal α-Mn2O3 for Catalytic Phenol Degradation in Aqueous Solution. Appl. Catal. B Environ. 2014, 154–155, 246–251. [Google Scholar] [CrossRef]
  56. Cheng, L.; Men, Y.; Wang, J.; Wang, H.; An, W.; Wang, Y.; Duan, Z.; Liu, J. Crystal Facet-Dependent Reactivity of α-Mn2O3 Microcrystalline Catalyst for Soot Combustion. Appl. Catal. B Environ. 2017, 204, 374–384. [Google Scholar] [CrossRef]
  57. Ji, F.; Men, Y.; Wang, J.; Sun, Y.; Wang, Z.; Zhao, B.; Tao, X.; Xu, G. Promoting Diesel Soot Combustion Efficiency by Tailoring the Shapes and Crystal Facets of Nanoscale Mn3O4. Appl. Catal. B Environ. 2019, 242, 227–237. [Google Scholar] [CrossRef]
  58. Liu, J.; Jiang, L.; Zhang, T.; Jin, J.; Yuan, L.; Sun, G. Activating Mn3O4 by Morphology Tailoring for Oxygen Reduction Reaction. Electrochim. Acta 2016, 205, 38–44. [Google Scholar] [CrossRef] [Green Version]
  59. He, D.; Li, Y.; Lyu, C.; Song, L.; Feng, W.; Zhang, S. New Insights into MnOOH/Peroxymonosulfate System for Catalytic Oxidation of 2,4-Dichlorophenol: Morphology Dependence and Mechanisms. Chemosphere 2020, 255, 126961. [Google Scholar] [CrossRef] [PubMed]
  60. Yan, H.; Shen, Q.; Sun, Y.; Zhao, S.; Lu, R.; Gong, M.; Liu, Y.; Zhou, X.; Jin, X.; Feng, X.; et al. Tailoring Facets of α-Mn2O3 Microcrystalline Catalysts for Enhanced Selective Oxidation of Glycerol to Glycolic Acid. ACS Catal. 2021, 11, 6371–6383. [Google Scholar] [CrossRef]
  61. Fan, Z.; Wang, Z.; Shi, J.-W.; Gao, C.; Gao, G.; Wang, B.; Wang, Y.; Chen, X.; He, C.; Niu, C. Charge-Redistribution-Induced New Active Sites on (0 0 1) Facets of α-Mn2O3 for Significantly Enhanced Selective Catalytic Reduction of NOx by NH3. J. Catal. 2019, 370, 30–37. [Google Scholar] [CrossRef]
  62. Wang, F.; Xiao, M.; Ma, X.; Wu, S.; Ge, M.; Yu, X. Insights into the Transformations of Mn Species for Peroxymonosulfate Activation by Tuning the Mn3O4 Shapes. Chem. Eng. J. 2021, 404, 127097. [Google Scholar] [CrossRef]
  63. Zhang, H.; Wang, X.; Li, Y.; Zuo, K.; Lyu, C. A Novel MnOOH Coated Nylon Membrane for Efficient Removal of 2,4-Dichlorophenol through Peroxymonosulfate Activation. J. Hazard. Mater. 2021, 414, 125526. [Google Scholar] [CrossRef] [PubMed]
  64. Shokoohi, R.; Khazaei, M.; Godini, K.; Azarian, G.; Latifi, Z.; Javadimanesh, L.; Zolghadr Nasab, H. Degradation and Mineralization of Methylene Blue Dye by Peroxymonosulfate/Mn3O4 Nanoparticles Using Central Composite Design: Kinetic Study. Inorg. Chem. Commun. 2021, 127, 108501. [Google Scholar] [CrossRef]
  65. Li, Y.; Li, D.; Fan, S.; Yang, T.; Zhou, Q. Facile Template Synthesis of Dumbbell-like Mn2O3 with Oxygen Vacancies for Efficient Degradation of Organic Pollutants by Activating Peroxymonosulfate. Catal. Sci. Technol. 2020, 10, 864–875. [Google Scholar] [CrossRef]
  66. Wang, Q.; Li, Y.; Shen, Z.; Liu, X.; Jiang, C. Facile Synthesis of Three-Dimensional Mn3O4 Hierarchical Microstructures for Efficient Catalytic Phenol Oxidation with Peroxymonosulfate. Appl. Surf. Sci. 2019, 495, 143568. [Google Scholar] [CrossRef]
  67. Zhang, L.; Tong, T.; Wang, N.; Ma, W.; Sun, B.; Chu, J.; Lin, K.A.; Du, Y. Facile Synthesis of Yolk–Shell Mn3O4 Microspheres as a High-Performance Peroxymonosulfate Activator for Bisphenol A Degradation. Ind. Eng. Chem. Res. 2019, 58, 21304–21311. [Google Scholar] [CrossRef]
  68. Zhao, Z.; Zhao, J.; Yang, C. Efficient Removal of Ciprofloxacin by Peroxymonosulfate/Mn3O4-MnO2 Catalytic Oxidation System. Chem. Eng. J. 2017, 327, 481–489. [Google Scholar] [CrossRef]
  69. Hu, L.; Deng, G.; Lu, W.; Lu, Y.; Zhang, Y. Peroxymonosulfate Activation by Mn3O4/Metal-Organic Framework for Degradation of Refractory Aqueous Organic Pollutant Rhodamine B. Chinese J. Catal. 2017, 38, 1360–1372. [Google Scholar] [CrossRef]
  70. Shokoohi, R.; Foroughi, M.; Latifi, Z.; Goljani, H.; Ansari, A.; Samarghandi, M.R.; Nematollahi, D. Comparing the Performance of the Peroxymonosulfate/Mn3O4 and Three-Dimensional Electrochemical Processes for Methylene Blue Removal from Aqueous Solutions: Kinetic Studies. Colloid Interface Sci. Commun. 2021, 42, 100394. [Google Scholar] [CrossRef]
  71. 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]
  72. Yermakov, A.N.; Zhitomirsky, B.M.; Poskrebyshev, G.A.; Sozurakov, D.M. The Branching Ratio of Peroxomonosulfate Radicals (SO5) Self-Reaction Aqueous Solution. J. Phys. Chem. 1993, 97, 10712–10714. [Google Scholar] [CrossRef]
  73. 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]
  74. Evans, D.F.; Upton, M.W. Studies on Singlet Oxygen in Aqueous Solution. Part 3. The Decomposition of Peroxy-Acids. J. Chem. Soc. Dalton Trans. 1985, 6, 1151–1153. [Google Scholar] [CrossRef]
  75. Ball, D.L.; Edwards, J.O. The Kinetics and Mechanism of the Decomposition of Caro’s Acid. I. J. Am. Chem. Soc. 1956, 78, 1125–1129. [Google Scholar] [CrossRef]
  76. Chen, C.; Xie, M.; Kong, L.; Lu, W.; Feng, Z.; Zhan, J. Mn3O4 Nanodots Loaded G-C3N4 Nanosheets for Catalytic Membrane Degradation of Organic Contaminants. J. Hazard. Mater. 2020, 390, 122146. [Google Scholar] [CrossRef]
  77. Chen, G.; Nengzi, L.-C.; Gao, Y.; Zhu, G.; Gou, J.; Cheng, X. Degradation of Tartrazine by Peroxymonosulfate through Magnetic Fe2O3/Mn2O3 Composites Activation. Chin. Chem. Lett. 2020, 31, 2730–2736. [Google Scholar] [CrossRef]
  78. Saputra, E.; Muhammad, S.; Sun, H.; Ang, H.-M.; Tadé, M.O.; Wang, S. A Comparative Study of Spinel Structured Mn3O4, Co3O4 and Fe3O4 Nanoparticles in Catalytic Oxidation of Phenolic Contaminants in Aqueous Solutions. J. Colloid Interface Sci. 2013, 407, 467–473. [Google Scholar] [CrossRef] [Green Version]
  79. Li, N.; Li, R.; Yu, Y.; Zhao, J.; Yan, B.; Chen, G. Efficient Degradation of Bentazone via Peroxymonosulfate Activation by 1D/2D γ-MnOOH-RGO under Simulated Sunlight: Performance and Mechanism Insight. Sci. Total Environ. 2020, 741, 140492. [Google Scholar] [CrossRef] [PubMed]
  80. Tian, N.; Tian, X.; Nie, Y.; Yang, C.; Zhou, Z.; Li, Y. Enhanced 2, 4-Dichlorophenol Degradation at PH 3–11 by Peroxymonosulfate via Controlling the Reactive Oxygen Species over Ce Substituted 3D Mn2O3. Chem. Eng. J. 2019, 355, 448–456. [Google Scholar] [CrossRef]
  81. Yao, Y.; Xu, C.; Yu, S.; Zhang, D.; Wang, S. Facile Synthesis of Mn3O4–Reduced Graphene Oxide Hybrids for Catalytic Decomposition of Aqueous Organics. Ind. Eng. Chem. Res. 2013, 52, 3637–3645. [Google Scholar] [CrossRef]
  82. Khan, A.; Zou, S.; Wang, T.; Ifthikar, J.; Jawad, A.; Liao, Z.; Shahzad, A.; Ngambia, A.; Chen, Z. Facile Synthesis of Yolk Shell Mn2O3@Mn5O8 as an Effective Catalyst for Peroxymonosulfate Activation. Phys. Chem. Chem. Phys. 2018, 20, 13909–13919. [Google Scholar] [CrossRef] [PubMed]
  83. Yang, B.; Tian, Z.; Wang, B.; Sun, Z.; Zhang, L.; Guo, Y.; Li, H.; Yan, S. Facile Synthesis of Fe3O4/Hierarchical-Mn3O4/Graphene Oxide as a Synergistic Catalyst for Activation of Peroxymonosulfate for Degradation of Organic Pollutants. RSC Adv. 2015, 5, 20674–20683. [Google Scholar] [CrossRef]
  84. Saputra, E.; Zhang, H.; Liu, Q.; Sun, H.; Wang, S. Egg-Shaped Core/Shell α-Mn2O3@α-MnO2 as Heterogeneous Catalysts for Decomposition of Phenolics in Aqueous Solutions. Chemosphere 2016, 159, 351–358. [Google Scholar] [CrossRef]
  85. Khan, A.; Wang, H.; Liu, Y.; Jawad, A.; Ifthikar, J.; Liao, Z.; Wang, T.; Chen, Z. Highly Efficient α-Mn2O3@α-MnO2-500 Nanocomposite for Peroxymonosulfate Activation: Comprehensive Investigation of Manganese Oxides. J. Mater. Chem. A 2018, 6, 1590–1600. [Google Scholar] [CrossRef]
  86. Huang, Y.; Nengzi, L.; Zhang, X.; Gou, J.; Gao, Y.; Zhu, G.; Cheng, Q.; Cheng, X. Catalytic Degradation of Ciprofloxacin by Magnetic CuS/Fe2O3/Mn2O3 Nanocomposite Activated Peroxymonosulfate: Influence Factors, Degradation Pathways and Reaction Mechanism. Chem. Eng. J. 2020, 388, 124274. [Google Scholar] [CrossRef]
  87. Shabanloo, A.; Salari, M.; Shabanloo, N.; Dehghani, M.H.; Pittman, C.U.; Mohan, D. Heterogeneous Persulfate Activation by Nano-Sized Mn3O4 to Degrade Furfural from Wastewater. J. Mol. Liq. 2020, 298, 112088. [Google Scholar] [CrossRef]
  88. Khan, A.; Zhang, K.; Sun, P.; Pan, H.; Cheng, Y.; Zhang, Y. High Performance of the A-Mn2O3 Nanocatalyst for Persulfate Activation: Degradation Process of Organic Contaminants via Singlet Oxygen. J. Colloid Interface Sci. 2021, 584, 885–899. [Google Scholar] [CrossRef] [PubMed]
  89. Li, Y.; Liu, L.-D.; Liu, L.; Liu, Y.; Zhang, H.-W.; Han, X. Efficient Oxidation of Phenol by Persulfate Using Manganite as a Catalyst. J. Mol. Catal. A Chem. 2016, 411, 264–271. [Google Scholar] [CrossRef]
  90. Xu, X.; Zhang, Y.; Zhou, S.; Huang, R.; Huang, S.; Kuang, H.; Zeng, X.; Zhao, S. Activation of Persulfate by MnOOH: Degradation of Organic Compounds by Nonradical Mechanism. Chemosphere 2021, 272, 129629. [Google Scholar] [CrossRef] [PubMed]
  91. Zhang, T.; Chen, Y.; Wang, Y.; Roux, J.L.; Yang, Y.; Croué, J.-P. An Efficient Peroxydisulfate Activation Process Not Relying on Sulfate Radical Generation for Water Pollutant Degradation. Environ. Sci. Technol. 2014, 48, 5868–5875. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  92. Wang, X.; Qin, Y.; Zhu, L.; Tang, H. Nitrogen-Doped Reduced Graphene Oxide as a Bifunctional Material for Removing Bisphenols: Synergistic Effect between Adsorption and Catalysis. Environ. Sci. Technol. 2015, 49, 6855–6864. [Google Scholar] [CrossRef] [PubMed]
  93. Lee, H.; Lee, H.-J.; Jeong, J.; Lee, J.; Park, N.-B.; Lee, C. Activation of Persulfates by Carbon Nanotubes: Oxidation of Organic Compounds by Nonradical Mechanism. Chem. Eng. J. 2015, 266, 28–33. [Google Scholar] [CrossRef]
  94. Ahn, Y.-Y.; Yun, E.-T.; Seo, J.-W.; Lee, C.; Kim, S.H.; Kim, J.-H.; Lee, J. Activation of Peroxymonosulfate by Surface-Loaded Noble Metal Nanoparticles for Oxidative Degradation of Organic Compounds. Environ. Sci. Technol. 2016, 50, 10187–10197. [Google Scholar] [CrossRef]
  95. Cheng, X.; Guo, H.; Zhang, Y.; Wu, X.; Liu, Y. Non-Photochemical Production of Singlet Oxygen via Activation of Persulfate by Carbon Nanotubes. Water Res. 2017, 113, 80–88. [Google Scholar] [CrossRef]
  96. Liu, Y.; Luo, J.; Tang, L.; Feng, C.; Wang, J.; Deng, Y.; Liu, H.; Yu, J.; Feng, H.; Wang, J. Origin of the Enhanced Reusability and Electron Transfer of the Carbon-Coated Mn3O4 Nanocube for Persulfate Activation. ACS Catal. 2020, 10, 14857–14870. [Google Scholar] [CrossRef]
  97. Rizal, M.Y.; Saleh, R.; Taufik, A.; Yin, S. Photocatalytic Decomposition of Methylene Blue by Persulfate-Assisted Ag/Mn3O4 and Ag/Mn3O4/Graphene Composites and the Inhibition Effect of Inorganic Ions. Environ. Nanotechnol. Monit. Manag. 2021, 15, 100408. [Google Scholar] [CrossRef]
  98. Ma, Q.; Zhang, X.; Guo, R.; Zhang, H.; Cheng, Q.; Xie, M.; Cheng, X. Persulfate Activation by Magnetic γ-Fe2O3/Mn3O4 Nanocomposites for Degradation of Organic Pollutants. Sep. Purif. Technol. 2019, 210, 335–342. [Google Scholar] [CrossRef]
  99. Yang, W.; Peng, Y.; Wang, Y.; Wang, Y.; Liu, H.; Su, Z.; Yang, W.; Chen, J.; Si, W.; Li, J. Controllable Redox-Induced in-Situ Growth of MnO2 over Mn2O3 for Toluene Oxidation: Active Heterostructure Interfaces. Appl. Catal. B Environ. 2020, 278, 119279. [Google Scholar] [CrossRef]
  100. Hazarika, K.K.; Talukdar, H.; Sudarsanam, P.; Bhargava, S.K.; Bharali, P. Highly Dispersed Mn2O3−Co3O4 Nanostructures on Carbon Matrix as Heterogeneous Fenton-like Catalyst. Appl. Organomet. Chem. 2020, 34, e5512. [Google Scholar] [CrossRef]
  101. Wang, Y.; Chen, L.; Cao, H.; Chi, Z.; Chen, C.; Duan, X.; Xie, Y.; Qi, F.; Song, W.; Liu, J.; et al. Role of Oxygen Vacancies and Mn Sites in Hierarchical Mn2O3/LaMnO3-δ Perovskite Composites for Aqueous Organic Pollutants Decontamination. Appl. Catal. B Environ. 2019, 245, 546–554. [Google Scholar] [CrossRef]
  102. Shokoohi, R.; Salari, M.; Shabanloo, A.; Shabanloo, N.; Marofi, S.; Faraji, H.; Tabar, M.V.; Moradnia, M. Catalytic Activation of Persulphate with Mn3O4 Nanoparticles for Degradation of Acid Blue 113: Process Optimisation and Degradation Pathway. Int. J. Environ. Anal. Chem. 2020, 1–20. [Google Scholar] [CrossRef]
  103. Liu, D.; Li, Q.; Hou, J.; Zhao, H. Mixed-Valent Manganese Oxide for Catalytic Oxidation of Orange II by Activation of Persulfate: Heterojunction Dependence and Mechanism. Catal. Sci. Technol. 2021, 11, 3715–3723. [Google Scholar] [CrossRef]
  104. Dong, Z.; Zhang, Q.; Chen, B.-Y.; Hong, J. Oxidation of Bisphenol A by Persulfate via Fe3O4-α-MnO2 Nanoflower-like Catalyst: Mechanism and Efficiency. Chem. Eng. J. 2019, 357, 337–347. [Google Scholar] [CrossRef]
  105. Buxton, G.V.; Greenstock, C.L.; Helman, W.P.; Ross, A.B. Critical Review of Rate Constants for Reactions of Hydrated Electrons, Hydrogen Atoms and Hydroxyl Radicals (⋅OH/⋅O) in Aqueous Solution. J. Phys. Chem. Ref. Data 1988, 17, 513–886. [Google Scholar] [CrossRef] [Green Version]
  106. 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]
  107. Kosmulski, M. Compilation of PZC and IEP of Sparingly Soluble Metal Oxides and Hydroxides from Literature. Adv. Colloid Interface Sci. 2009, 152, 14–25. [Google Scholar] [CrossRef] [PubMed]
  108. Guan, Y.-H.; Ma, J.; Li, X.-C.; Fang, J.-Y.; Chen, L.-W. Influence of PH on the Formation of Sulfate and Hydroxyl Radicals in the UV/Peroxymonosulfate System. Environ. Sci. Technol. 2011, 45, 9308–9314. [Google Scholar] [CrossRef]
  109. Chen, Z.; Li, X.; Zhang, S.; Jin, J.; Song, X.; Wang, X.; Tratnyek, P.G. Overlooked Role of Peroxides as Free Radical Precursors in Advanced Oxidation Processes. Environ. Sci. Technol. 2019, 53, 2054–2062. [Google Scholar] [CrossRef]
  110. He, D.; Niu, H.; He, S.; Mao, L.; Cai, Y.; Liang, Y. Strengthened Fenton Degradation of Phenol Catalyzed by Core/Shell Fe–Pd@C Nanocomposites Derived from Mechanochemically Synthesized Fe-Metal Organic Frameworks. Water Res. 2019, 162, 151–160. [Google Scholar] [CrossRef]
  111. Wang, L.; Jiang, J.; Pang, S.-Y.; Zhou, Y.; Li, J.; Sun, S.; Gao, Y.; Jiang, C. Oxidation of Bisphenol A by Nonradical Activation of Peroxymonosulfate in the Presence of Amorphous Manganese Dioxide. Chem. Eng. J. 2018, 352, 1004–1013. [Google Scholar] [CrossRef]
  112. Sun, S.P.; Hatton, T.A.; Chung, T.-S. Hyperbranched Polyethyleneimine Induced Cross-Linking of Polyamide−imide Nanofiltration Hollow Fiber Membranes for Effective Removal of Ciprofloxacin. Environ. Sci. Technol. 2011, 45, 4003–4009. [Google Scholar] [CrossRef] [PubMed]
  113. Du, X.; Zhang, Y.; Hussain, I.; Huang, S.; Huang, W. Insight into Reactive Oxygen Species in Persulfate Activation with Copper Oxide: Activated Persulfate and Trace Radicals. Chem. Eng. J. 2017, 313, 1023–1032. [Google Scholar] [CrossRef]
  114. Shih, Y.; Su, Y.; Ho, R.; Su, P.; Yang, C. Distinctive Sorption Mechanisms of 4-Chlorophenol with Black Carbons as Elucidated by Different PH. Sci. Total Environ. 2012, 433, 523–529. [Google Scholar] [CrossRef]
  115. Shah, N.S.; Ali Khan, J.; Sayed, M.; Ul Haq Khan, Z.; Sajid Ali, H.; Murtaza, B.; Khan, H.M.; Imran, M.; Muhammad, N. Hydroxyl and Sulfate Radical Mediated Degradation of Ciprofloxacin Using Nano Zerovalent Manganese Catalyzed S2O82−. Chem. Eng. J. 2019, 356, 199–209. [Google Scholar] [CrossRef]
  116. Huie, R.E.; Clifton, C.L.; Neta, P. Electron Transfer Reaction Rates and Equilibria of the Carbonate and Sulfate Radical Anions. Int. J. Radiat. Appl. Instrum. C Radiat. Phys. Chem. 1991, 38, 477–481. [Google Scholar] [CrossRef]
  117. Chen, S.-N.; Hoffman, M.Z.; Parsons, G.H. Reactivity of the Carbonate Radical toward Aromatic Compounds in Aqueous Solution. J. Phys. Chem. 1975, 79, 1911–1912. [Google Scholar] [CrossRef]
  118. Jiang, M.; Lu, J.; Ji, Y.; Kong, D. Bicarbonate-Activated Persulfate Oxidation of Acetaminophen. Water Res. 2017, 116, 324–331. [Google Scholar] [CrossRef] [Green Version]
  119. Fang, G.-D.; Dionysiou, D.D.; Wang, Y.; Al-Abed, S.R.; Zhou, D.-M. Sulfate Radical-Based Degradation of Polychlorinated Biphenyls: Effects of Chloride Ion and Reaction Kinetics. J. Hazard. Mater. 2012, 227–228, 394–401. [Google Scholar] [CrossRef]
  120. Qi, C.; Liu, X.; Li, Y.; Lin, C.; Ma, J.; Li, X.; Zhang, H. Enhanced Degradation of Organic Contaminants in Water by Peroxydisulfate Coupled with Bisulfite. J. Hazard. Mater. 2017, 328, 98–107. [Google Scholar] [CrossRef]
  121. Lei, Y.; Chen, C.-S.; Tu, Y.-J.; Huang, Y.-H.; Zhang, H. Heterogeneous Degradation of Organic Pollutants by Persulfate Activated by CuO-Fe3O4: Mechanism, Stability, and Effects of PH and Bicarbonate Ions. Environ. Sci. Technol. 2015, 49, 6838–6845. [Google Scholar] [CrossRef] [PubMed]
  122. Liang, C.; Wang, Z.-S.; Mohanty, N. Influences of Carbonate and Chloride Ions on Persulfate Oxidation of Trichloroethylene at 20 °C. Sci. Total Environ. 2006, 370, 271–277. [Google Scholar] [CrossRef]
  123. Zhou, J.; Xiao, J.; Xiao, D.; Guo, Y.; Fang, C.; Lou, X.; Wang, Z.; Liu, J. Transformations of Chloro and Nitro Groups during the Peroxymonosulfate-Based Oxidation of 4-Chloro-2-Nitrophenol. Chemosphere 2015, 134, 446–451. [Google Scholar] [CrossRef] [PubMed]
  124. Yang, Y.; Pignatello, J.J.; Ma, J.; Mitch, W.A. Comparison of Halide Impacts on the Efficiency of Contaminant Degradation by Sulfate and Hydroxyl Radical-Based Advanced Oxidation Processes (AOPs). Environ. Sci. Technol. 2014, 48, 2344–2351. [Google Scholar] [CrossRef] [PubMed]
  125. Grebel, J.E.; Pignatello, J.J.; Mitch, W.A. Effect of Halide Ions and Carbonates on Organic Contaminant Degradation by Hydroxyl Radical-Based Advanced Oxidation Processes in Saline Waters. Environ. Sci. Technol. 2010, 44, 6822–6828. [Google Scholar] [CrossRef] [PubMed]
  126. Yang, Y.; Pignatello, J.J.; Ma, J.; Mitch, W.A. Effect of Matrix Components on UV/H2O2 and UV/S2O82− Advanced Oxidation Processes for Trace Organic Degradation in Reverse Osmosis Brines from Municipal Wastewater Reuse Facilities. Water Res. 2016, 89, 192–200. [Google Scholar] [CrossRef] [Green Version]
  127. Minakata, D.; Kamath, D.; Maetzold, S. Mechanistic Insight into the Reactivity of Chlorine-Derived Radicals in the Aqueous-Phase UV–Chlorine Advanced Oxidation Process: Quantum Mechanical Calculations. Environ. Sci. Technol. 2017, 51, 6918–6926. [Google Scholar] [CrossRef]
  128. Wang, J.; Wang, S. Effect of Inorganic Anions on the Performance of Advanced Oxidation Processes for Degradation of Organic Contaminants. Chem. Eng. J. 2021, 411, 128392. [Google Scholar] [CrossRef]
Figure 1. Structural representations of α-Mn2O3 (a), γ-MnOOH (b), and (c) Mn3O4. The red, blue, and white balls represent oxygen, manganese, and hydrogen atoms, respectively. The black dashed lines represent the single unit cell. The crystalline parameters of Mn(III) (oxyhydr)oxides were taken from the crystallography open database (COD), and the COD ID of α- Mn2O3, γ-MnOOH, and Mn3O4 are 2105791, 1011012, and 1514121, separately [52,53,54].
Figure 1. Structural representations of α-Mn2O3 (a), γ-MnOOH (b), and (c) Mn3O4. The red, blue, and white balls represent oxygen, manganese, and hydrogen atoms, respectively. The black dashed lines represent the single unit cell. The crystalline parameters of Mn(III) (oxyhydr)oxides were taken from the crystallography open database (COD), and the COD ID of α- Mn2O3, γ-MnOOH, and Mn3O4 are 2105791, 1011012, and 1514121, separately [52,53,54].
Molecules 26 05748 g001aMolecules 26 05748 g001b
Figure 2. The activation mechanisms of peroxymonosulfate by Mn(III) (oxyhydr)oxides.
Figure 2. The activation mechanisms of peroxymonosulfate by Mn(III) (oxyhydr)oxides.
Molecules 26 05748 g002
Figure 3. The activation mechanisms of peroxydisulfate by various Mn(III) (oxyhydr)oxides.
Figure 3. The activation mechanisms of peroxydisulfate by various Mn(III) (oxyhydr)oxides.
Molecules 26 05748 g003
Figure 4. The diagram of PDS activation on the surface of γ-MnOOH. The red, blue, and white balls in the structure of γ-MnOOH represent the oxygen, manganese, and hydrogen atoms, respectively. The COD ID of γ-MnOOH is 1011012 [54].
Figure 4. The diagram of PDS activation on the surface of γ-MnOOH. The red, blue, and white balls in the structure of γ-MnOOH represent the oxygen, manganese, and hydrogen atoms, respectively. The COD ID of γ-MnOOH is 1011012 [54].
Molecules 26 05748 g004
Table 1. The structures of common Mn(III) (oxyhydr)oxides [50,51].
Table 1. The structures of common Mn(III) (oxyhydr)oxides [50,51].
Mineral NameChemical FormulaMn ValenceCrystal Structure
Mn(III) oxideα-Mn2O3IIIBixbyite
Groutiteα-MnOOHIIITunnel
Feitknechtiteꞵ-MnOOHIIILayer
Manganiteγ-MnOOHIIITunnel
HausmanniteMn3O4II/IIISpinel
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Jia, D.; Hanna, K.; Mailhot, G.; Brigante, M. A Review of Manganese(III) (Oxyhydr)Oxides Use in Advanced Oxidation Processes. Molecules 2021, 26, 5748. https://doi.org/10.3390/molecules26195748

AMA Style

Jia D, Hanna K, Mailhot G, Brigante M. A Review of Manganese(III) (Oxyhydr)Oxides Use in Advanced Oxidation Processes. Molecules. 2021; 26(19):5748. https://doi.org/10.3390/molecules26195748

Chicago/Turabian Style

Jia, Daqing, Khalil Hanna, Gilles Mailhot, and Marcello Brigante. 2021. "A Review of Manganese(III) (Oxyhydr)Oxides Use in Advanced Oxidation Processes" Molecules 26, no. 19: 5748. https://doi.org/10.3390/molecules26195748

Article Metrics

Back to TopTop