Next Article in Journal
Facile and Novel Route for the Preparation of ZnO Nanoparticles with Different Cr Loadings for Opto-Photocatalysis Applications
Previous Article in Journal
Waste Biomass Selective and Sustainable Photooxidation to High-Added-Value Products: A Review
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

A Review of Sulfate Radical-Based and Singlet Oxygen-Based Advanced Oxidation Technologies: Recent Advances and Prospects

1
Research Centre for Coastal and Marine Resource Utilization, Tianjin Research Institute for Water Transport Engineering, Ministry of Transport, Tianjin 300456, China
2
Construction Management Department, Tianjin Eco-City Water Investment and Construction Co., Ltd., Tianjin 300457, China
3
Faculty of Environmental Science and Engineering, Nankai University, Tianjin 300350, China
*
Author to whom correspondence should be addressed.
Catalysts 2022, 12(10), 1092; https://doi.org/10.3390/catal12101092
Submission received: 5 September 2022 / Revised: 16 September 2022 / Accepted: 17 September 2022 / Published: 21 September 2022

Abstract

:
In recent years, advanced oxidation process (AOPs) based on sulfate radical (SO4●−) and singlet oxygen (1O2) has attracted a lot of attention because of its characteristics of rapid reaction, efficient treatment, safety and stability, and easy operation. SO4●− and 1O2 mainly comes from the activation reaction of peroxymonosulfate (PMS) or persulfate (PS), which represent the oxidation reactions involving radicals and non-radicals, respectively. The degradation effects of target pollutants will be different due to the type of oxidant, reaction system, activation methods, operating conditions, and other factors. In this paper, according to the characteristics of PMS and PS, the activation methods and mechanisms in these oxidation processes, respectively dominated by SO4●− and 1O2, are systematically introduced. The research progress of PMS and PS activation for the degradation of organic pollutants in recent years is reviewed, and the existing problems and future research directions are pointed out. It is expected to provide ideas for further research and practical application of advanced oxidation processes dominated by SO4●− and 1O2.

1. Introduction

With the development of economy and science and technology, more and more new substances are developed and manufactured to meet the growing needs of human beings. At the same time, many emerging pollutants, such as pesticides, pharmaceuticals and personal care products (PPCPs), polycyclic aromatic hydrocarbons (PAHs), perfluorinated compounds (PFCs), food additives, surfactants, plasticizers, and pathogens, are released into the aquatic environment during production, processing, use, and waste. As these pollutants migrate and transform in the environment, they accumulate in organisms and eventually harm human health. These emerging pollutants have relatively stable properties, complex molecular structures and are difficult to degrade in the natural environment, so they cannot be completely degraded by traditional biological processes [1,2].
In recent decades, water treatment technologies have achieved rapid development. Especially, advanced oxidation processes (AOPs) have been applied to the degradation, complete mineralization or pretreatment of complex and refractory organic pollutants. Compared with traditional water treatment methods, advanced oxidation technology has obvious advantages of strong oxidation capacity, fast reaction rate, simple operation, wide application range, high treatment efficiency, less secondary pollution, high mineralization rate, equipment, and large-scale application. AOPs relies on reactive oxygen species (ROS) to degrade organic pollutants. Among them, sulfate radical (SO4●−) and singlet oxygen (1O2) have attracted more and more researchers’ attention due to their unique properties in the removal of organic pollutants [3,4]. Compared with the traditional hydroxyl radical (OH), SO4●− has a higher redox potential (2.5–3.1 V), and the redox potential of OH is 2.7 V and 1.8 V under acidic and alkaline conditions, respectively. In addition, SO4●− (30–40 μs) has a longer half-life than OH (20 ns) [5]. SO4●− also degrades many organic compounds more efficiently than OH, due to its greater selectivity for electron transfer reactions, giving SO4●− a greater advantage in mineralizing a wide range of organic pollutants [6]. Moreover, the oxidation efficiency of SO4●− in carbonate and phosphate buffer solutions is higher than that of OH [7]. Therefore, sulfate radical-based AOPs will have better oxidation efficiency and application than hydroxyl radical-based AOPs.
Unlike SO4●− and OH, 1O2 is a non-free radical ROS, which can degrade organic pollutants through non-free radical attack pathways [8]. 1O2 is electrophilic and highly selective for electron-rich organic matter, which facilitates selective removal of micropollutants (e.g., drugs and endocrine disruptors) in the presence of salinity and other organic matter [9,10]. Therefore, the non-radical oxidation reaction based on 1O2 provides an idea for the removal of refractory organic compounds in complex matrix.
SO4●−-based and 1O2-based AOPs have attracted increasing interest. Most of the recent studies focus on the generation of SO4●− and 1O2 and the degradation of one or several organic pollutants. However, there are few reports on the comprehensive evaluation of the formation, mechanism, and application of SO4●− and 1O2. This paper reviews the recent progress and prospect about sulfate radicals-based and singlet oxygen-based advanced oxidation technologies. The production of SO4●− and 1O2 mainly comes from the activation of persulfate (PS) and peroxymonosulfate (PMS). The chemical properties of PMS and PS are briefly introduced in this paper, and the generation, reaction mechanism, and treatment effect of SO4●−-based and 1O2-based AOPs are emphasized. The obstacles in the current research work are pointed out and the future research is prospected. It is expected to provide ideas for further research and practical application of AOPs based on SO4●− and 1O2. This paper has a certain reference and guiding significance for the applications of PMS or PS activation technologies in environmental pollution control and remediation.

2. Characteristics of Persulfate and Peroxymonosulfate

At present, the production of SO4●− mainly comes from the activation of PMS and PS, which are the monosubstituted or symmetrically substituted derivatives of hydrogen peroxide by sulfonic acid group (-SO3), respectively. PMS has been widely used in organic compound synthesis and as a chlorine-free additive for disinfecting swimming pools at a rate of about 1–2 pounds per 10,000 gallons of pool water [11]. Under the presence of 25 mg/L PMS and 0.1 mg/L Co2+, the removal rate of E. coli reached 99.99% after 1 h of reaction [12]. PMS is white solid powder. It is stable when pH is less than 6 or pH is 12. When pH is 9, it showed the poorest stability where half of HSO5 decomposes to SO52− [4]. At present, the widely used potassium bisulfate complex salt (2KHSO5·KHSO4·K2SO4) is composed of three components, potassium peroxymonosulfate, potassium hydrogen sulfate, and potassium sulphate, and its main active substance is potassium peroxymonosulfate (KHSO5). The salt is marketed under the trade names Caroat and Oxone registered by Evonik and DuPont, respectively. Oxone is a white granular powder crystal salt, which is stable, non-toxic, inexpensive, and soluble in water. The peroxide bond (O-O) distance is 1.453 Å, and the bond energy is 140–213.3 kJ/mol. PMS is most stable when the solution pH is less than 6 and equal to 12. When pH is 9, the stability is worst, and the concentration of HSO5 and SO52− in the solution is almost equal. When pH is less than 1, PMS will undergo hydrolysis reaction to produce H2O2 [13]. PS, an oxidant with symmetrical structure, was first known as the initiator of polymerization reaction. The O-O distance is 1.497 Å, and the bond energy is 140 kJ/mol [14]. PS, often in the form of potassium persulfate or sodium persulfate, has been widely used as bleaching agents, oxidants, emulsion polymerization promoters, and water or soil remediation agents. The related properties of PMS and PS are shown in Table 1. Both PMS and PS are strong oxidants, but their direct reaction rates with most pollutants are very low. Therefore, it is necessary to activate them through appropriate ways to destroy the O-O bond and generate strong oxidizing free radicals, 1O2 and other ROS to degrade organic pollutants quickly and efficiently.
Because PS and PMS are solid powders, they can be transported and stored more easily. Compared with H2O2, the anions of PS and PMS remain stable in water for a much longer time until they are properly activated. In addition, PMS and PS-based AOPs can proceed smoothly in a wide solution pH range from acidic to alkaline (pH = 2–10), while H2O2-based Fenton process requires strict acidic conditions (pH = 2.7–3). Generally, PS and PMS can be activated with the assistance of ultraviolet light, heat, alkali, or metal catalysts, etc. Different types of oxidants and activation methods will produce different ROS. The advanced oxidation processes dominated by SO4●− and 1O2 were discussed in this paper.

3. Sulfate Radicals-Based Advanced Oxidation

Studies have shown that AOPs based on SO4●− and OH is an effective method for the degradation of refractory organic pollutants, such as pharmaceuticals, pesticides, personal care products, steroids, endocrine disruptors, etc. [15]. However, AOPs based on OH degrades organic pollutants through a non-selective, multi-step approach that typically requires an acidic environment. In addition, the oxidation process is severely limited by the large amount of dissolved organic matter and anions in complex environments, which are the main scavengers of OH. SO4●− is inherently more oxidizing than OH and lasts longer in aqueous solutions, and in some cases SO4●− can oxidize contaminants that OH cannot. In recent years, SO4●−-based AOPs have replaced OH-based AOPs to some extent.
The essence of the advanced oxidation process based on SO4●− is to activate PMS or PS to form SO4●− to achieve the removal of organic matter. Besides SO4●−, there are also associated or indirect generation of OH, superoxide radical (O2●−), or other radicals, but SO4●− plays a leading role in the degradation of pollutants. These free radicals (SO4●−, OH, and O2●−) could be detected by using 5,5-dimethyl-1-pyrroline-1- oxide (DMPO) as a spin trapping agent in an electron paramagnetic resonance spectroscopy (EPR) [16]. Both ethanol (Et) and tert-butanol (TBA) could quench OH rapidly (kEt = 1.2–2.8 × 109 M−1s−1, kTBA = 3.8–7.6 × 108 M−1s−1), and the reaction rate between Et and SO4●− is much faster than that of TBA (kEt = 1.6–7.7 × 107 M−1s−1, kTBA = 4.0–9.1 × 105 M−1s−1). Therefore, Et and TBA can be used as capture agents to identify who contributes more to the degradation of pollutants [17]. PMS and PS can be activated to produce SO4●− through energy input, transition metal ions and their oxides, non-metallic materials, etc. Studies on the degradation of pollutants by activating PMS and PS in different ways to generate SO4●− are summarized in Table 2.

3.1. Energy Activation

The activation process of PMS and PS through energy is shown in Figure 1. Energy input methods include heating, ultraviolet light, electricity, ultrasound, and radiation. It has been reported that at least 33.5 kcal/mol is required to break the O-O bond of S2O82 by heat [46]. Therefore, PS activation occurs only when the temperature reaches the required activation energy. Liang et al. found that PS could not be activated to achieve degradation of trichloroethylene and 1,1, 1-trichloroethane at 20 °C, whereas pollutants degradation could occur at 40–60 °C, and they found that the reaction with higher activation energy was more sensitive to temperature [47,48]. Yang et al. found that heating activation of PS is effective, but the activation efficiency of PMS is very low [49]. The combination of ultrasound and light with PS or PMS has received much attention in recent years as a potential alternative method for removal of refractory or toxic compounds. When H2O2, PS and PMS were activated by UV light, the removal rate of was 100% in the presence of PS and PMS, but only 75% in the presence of H2O2 [50]. 1,1, 1-trichloroethane could be completely removed by ultrasonic activation of sodium persulfate, and the main ROS were SO4●− and OH [23]. It has also been reported that laser can activate PS to produce SO4●−, but there are few studies on its application in pollutant removal in recent years [51]. Although energy input can activate PMS and PS to produce free radicals, the generation efficiency of free radicals is generally low, accompanied by large energy output and high cost.

3.2. Transition Metal Ions Activation

In addition to energy activation, transition metal ions such as Co2+, Ce3+, Mn2+, Ni2+, and Fe2+can also activate PMS and PS. The reaction processes are redox reaction, as shown in Equations (1) and (2) [46]. There are many studies on the activation of PMS by Co2+, and the relevant mechanism is basically clear, as shown in Figure 2. It mainly converts Co2+ into CoOH+, then generates SO4●− radical and CoO+, further converts into Co3+, and finally turns into Co2+ to complete the catalytic process [52]. The results show that Co2+ is the most efficient metal for activating PMS, while Ag+ is the most efficient metal for activating PS [30]. Co2+/PMS system is better than traditional Fenton oxidation, which can not only react under neutral conditions, but also has lower drug consumption [30]. CoOH+ is the most effective intermediate substance in the reaction process of Co2+ activation of PMS, and its formation is the rate-limiting step of the reaction, as shown in Equation (3). The regeneration of Co2+ is realized through the reduction reaction of Co3+, which is a key step to maintain the low usage of cobalt, as shown in Equation (4) [5].
S2O82− + Mn → Mn+1 + SO4●− + SO42−
HSO5+ Mn → Mn+1 + SO4●− + OH
CoOH+ + HSO5 → CoO+ + SO4●− + H2O
Co3+ + HSO5 → Co2+ + SO5●− + H2O
Although cobalt ions have good activation ability of PMS, cobalt metals are toxic. The average concentration of cobalt in human serum and urine is about 0.1–0.3 μg/L and 0.1 μg/L, respectively [53]. Studies have shown that excessive cobalt may be potentially toxic and carcinogenic to human body, causing some diseases, such as asthma, pneumonia, and cardiomyopathy [5]. Therefore, the activation of PMS with cobalt-based materials has great environmental risks. Ag+ has better activation effect on PS, but silver is a precious metal, which is not easy to achieve large-scale application because of high cost. Iron-based materials have attracted widespread attention because of their non-toxic, large reserves and low price. Fe2+ can activate PMS and PS effectively, while Fe3+ has poor activation effect [54]. For homogeneous reaction systems, dissolved metal ions can react freely with PS and PMS, so mass transfer has little influence on the activation of PS and PMS. However, homogeneous reaction also has some disadvantages: (1) the dissolved metal ions are difficult to recover; (2) high concentration of organic wastewater will use a lot of metal salts, resulting in secondary pollution of metal ions in sewage; (3) transition metal ions are easily affected by water quality components and pH, such as precipitation under alkaline conditions, hydrolysis under acidic conditions, thus reducing the activation efficiency; and (4) in addition, some organic compounds in sewage will be complex with metal ions, which will also affect the activation of PS or PMS.

3.3. Transition Metal Oxide Activation

In order to overcome the problem of secondary contamination of homogeneous catalysts, researchers gradually shifted their research focus to heterogeneous catalytic reactions. Transition metal oxides also have good activation effect on persulfate. For example, five typical cobalt oxides (CoO, CoO2, Co2O3, CoO(OH), and Co3O4) have been reported. Among them, CoO and Co3O4 have been studied the most frequently, while CoO2 has thermal instability and receives less attention. Dionysiou’s group first studied the activation of PMS by CoO or Co3O4, and found that CoO/PMS systems tend to be homogeneous reactions because CoO has a high solubility in water (0.313 mg/100 g H2O) [7]. When Co3O4 was used to activate PMS, the amount of cobalt dissolved was significantly reduced, mainly because CoO was inserted into the lattice of Co2O3, which made the structure relatively stable [55]. Co3O4 has been widely studied for activating PMS due to its relatively high stability. In addition to the synthesis of Co3O4 nanomaterials with various morphs, researchers also composite it with various carriers to prepare a variety of cobalt-supported materials to activate PMS. Carbon fiber, activated carbon, graphene, red mud, zeolite, SBA-15, and other materials are used as support materials [13]. Potential environmental risks and high cost limit the application of cobalt materials. In order to replace cobalt-based materials, other metal oxides such as MnO2, CuO, Fe3O4, Fe2O3, ZnO, and MFe2O4 (M = Cu, Co, Mn, and Zn) have also been applied to PMS and PS activation studies, as shown in Figure 3. The mechanisms of PMS and PS activation by metal oxides are similar to that of homogeneous metal ions, which mainly depends on the redox action of metal ions [6]. According to the reduction mechanism, the activation of PS and PMS is positively correlated with the redox potential of the metal. In fact, metal redox potential is not the only factor affecting the activation of PS and PMS. For example, the redox potential of Ce3+ is higher than that of Fe2+, but the activation effect is worse than that of Fe2+ [56]. Therefore, the mechanism and process of metals activating PMS or PS still need to be further studied, and low-cost, efficient, and stable metal catalytic materials also need to be developed and applied.

3.4. Non-Metallic Materials Activation

Recently, the most studied non-metallic catalytic materials are carbon-based materials, which have been widely used as adsorbent or carrier materials due to their large specific surface area, low cost, and large pore capacity. It was found that carbon-based materials can also activate PMS and PS well, and electron conduction is the main mechanism. It was found that carbon-based materials can also activate persulfate well, and electron conduction is the main mechanism. For example, the sp2 covalent carbon net structure of multiwalled carbon nanotube forms a redox cycle with the oxygen-containing functional group at the edge defect, which provides electrons to activate PS to generate SO4●− radicals [57]. In another form, electrons may be transferred from the carbon-based material of the aromatic graphite structure to PMS or PS [58]. PS and PMS are activated to produce SO4●− radicals by receiving electrons, as shown in Equations (5) and (6). The structure and composition of carbon-based materials affect the redox potential and further affect the electron transfer reactions.
S2O82− + e → SO4●− + SO42−
HSO5+ e → SO4●− + OH
S2O82− + H2O → 2SO42− + HO2 + H+
S2O82− + HO2 → SO42− + SO4●− + O2●− + H+
S2O82− + PhO → SO42− + SO4●− + PhOOX
In addition, alkali, phenols, and quinones can also activate PS to generate SO4●− radicals, as shown in Equations (7)–(9), while they activate PMS to generate singlet oxygen [6]. In order to improve the activation efficiency, the catalyst will be added at the same time with UV light, ultrasound and other energy input to promote the generation of free radicals.
Heterogeneous catalysis technology has the characteristics of high efficiency, low energy consumption, continuous catalysis, catalyst recovery, and less secondary pollution, which makes it become the mainstream research direction in many studies and shows a better application prospect. PMS and PS heterogeneous activation relies only on catalysts and oxidants, so the key lies in the development of low-cost, stable and efficient heterogeneous catalysts. Current reports mainly focus on the synthesis of laboratory-scale catalysts and the degradation of emerging pollutants. The pollutants treated are relatively single, and there is a lack of data for actual wastewater treatment. In addition, the preparation of catalytic materials reported is relatively complex, which is still a long way from large-scale production.

4. Singlet Oxygen-Based Advanced Oxidation

With the report of PMS and PS non-radical activation, people gradually have a strong interest in the process of non-radical oxidation. Non-radical reactive species are generally thought to be resistant to common free radical scavengers (e.g., methanol, ethanol, and tert-butanol), selective to electron-rich organic compounds, and particularly sensitive to organic substrates with mild redox potentials [8]. At present, non-radical oxidation processes have been found in a variety of reaction systems, especially in PMS and PS activation. However, the mechanism remains controversial. Based on current reports, carbon or metal catalysts mainly realize non-radical activation of oxidants (PMS, PS, H2O2, or O3) through three ways: electron transfer process, generation of activated complexes (or surface-bound radicals) and singlet oxygen participating in pollutant degradation, respectively, as shown in Figure 4. Among them, 1O2 shows significant advantages in the selective removal of organic pollutants, which has become the focus of research.

4.1. Properties of Singlet Oxygen

1O2 is an excited molecular oxygen, which is the excited state of the ground state oxygen (triplet oxygen molecule, 3Σg). Originally, two electrons in two 2pπ* orbitals with two spins parallel to each other can occupy one 2pπ* orbital with opposite spins, or two 2pπ* orbitals with opposite spins [59]. They’re labeled as O2(1g) and O2 (1Σg+). Since O2 (1Σg+) has a short lifetime (10−12 s) and is easily converted to low-excited O2(1g) (10−3–10−6 s), what is commonly called singlet oxygen is O2(1g) [60]. 1O2 is a non-radical substance with an energy of 94.2 kJ/mol, higher than the ground state oxygen molecule (O2). As a mildly oxidizing species, the standard redox potential (E0 = 1.52 V) of 1O2 was significantly lower than OH and SO4•− [59]. Due to the vacant orbitals, 1O2 can obtain electrons from other substances and is a kind of electrophilic reagent with mild reaction. Therefore, it can quickly oxidize organic pollutants with electron-rich functional groups and is not easy to be inactivated by background components in water [61]. Due to unique properties, 1O2 has been used for selective oxidation [62], degradation of organic pollutants [10], inactivation of pathogenic bacteria [63], cancer treatment [64], etc. 1O2 could be detected by chemical probe tests, electron paramagnetic resonance (EPR) spectroscopy and chemiluminescence detection. 1O2 could rapidly react with sodium azide (k = 1 × 109 M−1s−1) and furfuryl alcohol (k = 1.2 × 108 M−1s−1), but it is insensitive to alcohols (such as Et and TBA) [65]. Moreover, 1O2 can be detected by using 2,2,6,6-tetramethyl-4-piperidinol (TEMP) as a spin trapping agent in EPR with an intensive 1:1:1 signal [16]. In addition, solid-state near infrared spectroscopy and Fluorescence detection are helpful technique for 1O2 detection [64,66].
Previous reports have indicated that 1O2 can be produced by a photochemical process that uses photosensitization to transfer light energy to molecular oxygen [67]. For example, Dorota et al. successfully inactivated Escherichia coli and Aspergillus fumigatus using a miniature photoreactor to produce 1O2 [68]. In artificial or natural water, natural organic matter is used as photosensitizer to generate 1O2 through sunlight irradiation [69,70]. Other studies have shown that 1O2 can be generated by photoactivation of oxygen molecules in the oxidation of semiconductor metals such as TiO2 [71]. Studies on the generation of 1O2 by non-photosensitive methods are also emerging. It has been reported that catalysts such as BiAgxOy and Bi(V)-Bi(III) complex were used to release lattice oxygen to generate 1O2, which could remove pollutants such as Rhodamine B and Bisphenol A [72,73]. Similarly, MgO with high-energy {111} crystal planes and α-Bi2O3 with oxygen vacancies can react with O2 to produce 1O2 [74]. In addition, phosphite reaction with ozone [75], H2O2 reaction with NaClO [76], and the activation of periodate acid [77] can generate 1O2. With the development of PMS and PS activation studies, 1O2 can also be generated by non-radical activation of PMS or PS, which provides a new path for the selective removal of organic pollutants. The relevant studies are summarized in Table 3.

4.2. Homogeneous Activation

In recent years, it has been reported that 1O2 is produced by activated PMS via non-radical pathway leading to organic pollutants degradation. It has been found that ketones, phenols, quinones, and other hydrocarbons can activate PMS to produce 1O2. Zhou et al. reported that benzoquinone can activate PMS to produce 1O2 through a non-free radical pathway, and found that the removal rate of sulfamethoxazole gradually increased with the increase in solution pH from 7 to 10 [83]. As early as 1974, it was reported that ketones can activate PMS and that the decomposition rate of PMS is proportional to the amount of low concentration ketones [96]. Later, Lange and Brauer detected 1O2 at 1270 nm using infrared phosphorescence measurement technology, confirming the formation of 1O2 in the ketone-catalyzed PMS process [97]. Since phenolic compounds are easily oxidized to quinone as by-products, phenols can also activate PMS. Zhou et al. found that PMS was activated by phenol to generate 1O2 at pH 8.5 and 10, and phenol was oxidized to benzoquinone to further promote the activation reaction [55]. However, under acidic conditions, the dissociation of phenol is poor and it is difficult to form intramolecular complex with ionized PMS, resulting in difficult activation reaction
In fact, alkali has been used as a typical PS activator in site restoration, while it activates PMS to produce 1O2 [79]. PMS can spontaneously decompose to produce 1O2 under alkaline conditions, and the reaction rate constant is about 0.2 M−1s−1 [78]. Moreover, polyphosphate [80], phosphite (HPO32−) [81], BO2 [82], and CO32− [83] have also been found to activate PMS to produce 1O2 to degrade organic pollutants. All of the above are homogeneous reactions, activators are added into solution continuously to ensure continuity of the reaction, which is prone to high costs and secondary pollution. Therefore, heterogeneous activation reaction will be a good solution.

4.3. Heterogeneous Activation

Metal oxides and complexes can also activate PMS and PS in a non-radical way. Bu et al. found that oxygen vacancy on bismuth bromide (BiOBr) could activate PS to produce 1O2 as the main active species to degrade Bisphenol A under alkaline conditions [98]. The sillenite Bi25FeO40 was synthesized by hydrothermal method and used to activate PMS to degrade levofloxacin [99]. The results showed that 1O2 was the main ROS. Yang et al. synthesized Fe0-montmorillonite composite material (Fe-Mt-C-H2) to activate PMS for the degradation of Bisphenol A [100]. It could remove 99.3% of Bisphenol A (25 mg/L) at pH 3 with 0.4 g/L Fe-Mt-C-H2 and 1 mM PMS. The mineralization rate was 70.6%, and the degradation of pollutants was dominated by 1O2 and superoxide radical (O2●−). Liu et al. used Co3O4-SnO2/rice straw biochar (0.1 g/L) to activate PMS (1 mmol/L) to degrade sulfamoxole (50 mg/L), which was degraded 98% in 5 min by 1O2 [101]. CoOOH [102] and γ-MnOOH [103] were also used to activate PMS to degrade 2,4-dichlorophenol, which proved that 1O2 had a good removal effect on it. Li et al. used CuO-based modified materials to activate PMS to treat high-salt organic wastewater, and confirmed that the non-radical oxidation process led by 1O2 has a good removal effect on heterocyclic organic pollutants in a complex background [10,87,94]. Although the catalysis of PS or PMS by various metal oxides and their complexes has been studied endlessly, the environmental risks caused by metal ion leaching still exist in metal catalysts, and the separation and reuse of catalysts also face challenges.
A variety of carbon materials, including carbon nanotubes (CNTs), graphene oxide (GO), reduced graphene oxide (rGO), carbon spheres, nano-diamond and biochar, have been demonstrated to activate persulfate through a 1O2-dominated non-radical pathway [104]. Cheng et al. used CNTs and PS combined system to efficiently generate 1O2 under no irradiation, which could selectively oxidize 2,4-dichlorophenol in water under neutral conditions [61]. CNTs not only act as electronic bridges between PS and organic pollutants, but also play an important role in the presence of various oxygen-containing functional groups (such as -COOH, C=O, and -OH) on the surface of CNTs, because they directly affect the zeta potential of CNTs, which is conducive to the absorption of PS, and promote the non-radical oxidation process of target pollutants [105]. GO and its reduced derivatives rGO show great potential in hydrophilic adsorption of organic matter and PMS/PS catalysis due to various structural defects, vacancies and oxygen-containing functional groups. Density Functional theory (DFT) calculations show that the vacancy and defect edges of rGO elongate the O-O bond of PMS, enhance adsorption and direct electron transfer, and promote the final rupture of the O-O bond, leading to non-radical oxidation [106]. The increased degree of carbonyl and graphitization results in more vacancies and defects, thus improving the catalytic performance of GO and rGO [107]. N-doping is also an effective strategy. N-doped carbon is expected to have more lattice defects to regulate the electronic structure, such as sp2 hybrid carbon skeleton [108]. Doping nitrogen into the carbon matrix not only facilitates the adsorption of PMS, increases the surface alkalinity, but also facilitates the transfer of electrons to the negatively charged PMS, thus improving the catalytic activity. Biomass type and pyrolysis temperature are important factors affecting the structure and catalytic activity of biochar. Similarly, more oxygen vacancies, defects and oxygen-containing functional groups are still important factors in achieving 1O2 production from non-radical activated PMS and PS [109]. Carbon materials can not only achieve catalytic function, but also have the ability to adsorb pollutants, so it has great application potential. However, carbon materials have a variety of structural and surface characteristics, and the identification of their internal active sites is still difficult and controversial.
The studies on the degradation of organic pollutants by 1O2 have been reported, but there are still many problems to be solved, such as the preparation of high efficiency and stable catalysts, optimization of catalyst scale production process, clarifying the activation mechanism from different catalysts, explaining the interaction processes of 1O2 with organic pollutants, and effect of coexisting substances in wastewater on organic degradation process. In particular, few studies have been reported on the removal of organic pollutants in high-salt wastewater. The removal effect of organic pollutants in high-salt system, degradation kinetics, degradation intermediates and toxicity, the influence of inorganic salts type and concentration on the degradation process, and catalytic reaction mechanism remain to be solved.

5. Conclusions and Prospect

In this paper, the related research on SO4●−-based and 1O2-based AOPs has been reviewed, and the production methods and mechanism of SO4●− and 1O2 are analyzed. Compared with the advanced oxidation process dominated by OH, the SO4●−-based AOPs have a series of advantages such as long half-life of SO4●−, wide range of pH application and strong oxidation capacity, which have a greater advantage in the removal of refractory pollutants. The non-radical oxidation process based on 1O2 has certain resistance to a variety of inorganic ions, halogens and background organic matter, strong selectivity and moderate redox potential, which can be used for the pre-oxidation treatment of highly dangerous industrial wastewater, pharmaceutical and pesticide wastewater. The SO4●−-based and 1O2-based AOPs have their own characteristics, which have good application prospects in the field of water purification and the removal of refractory organic pollutants from complex matrix. According to the existing research reports, SO4●−-based and 1O2-based AOPs still have some limitations.
(1) Most existing research has focused on treating single-component pollutants in the laboratory. However, actual wastewater is complex and variable. Although many research studies have been very successful on a laboratory scale, many problems need to be solved before large-scale practical applications. Especially, it should pay attention to addressing amplification issues such as design challenges, low energy requirements, and complex water quality changes. The next step is to study multi-component organic wastewater in real environment, and systematically study the reaction mechanisms by combining various characterization techniques and theoretical calculations, so as to further promote the popularization and application of this technology.
(2) There are still controversial including the mechanisms of PMS and PS activated by various catalysts. For example, the biochar-based materials will generate a variety of morphological structures and surface functional groups due to different raw materials and preparation methods. Therefore, the mechanisms of activating PMS and PS will be different. It is necessary to systematically study the relationship between material configuration and the efficiency of SO4●− and 1O2 production. In addition, the catalysts should be prepared from materials that are easily available and cheap, and the preparation process should be simple, energy saving, and easy to large-scale production.
(3) There are some unavoidable drawbacks for a single activation method. For example, thermal activation will lead to higher energy consumption, alkaline activation will affect the pH value of the system, electrochemical activation will produce more anode slime, carbon-based catalyst activation will make it difficult to regenerate, homogeneous transition metal activation will produce metal ions and sludge, resulting in secondary pollution, etc. The next step is to overcome the shortcomings of each activation method and maximize the development of multiple combined activation methods.
(4) Considering that 1O2 has strong anti-interference ability to inorganic salts and natural organics, while SO4 and OH have strong oxidation ability, the reaction system can be designed according to the treated objects to achieve phased and multi-channel activation of PS and PMS. It is expected to achieve higher oxidation efficiency by selectively regulating the production of strong oxidizing radicals (SO4 and OH) and non-radical (1O2) according to water quality and pollutant types and concentration levels, which will provide ideas for the practical application of this technology.
(5) In addition to the removal of refractory pollutants, 1O2 as a non-radical ROS can be used to inactivate pathogenic microorganisms and remove antibiotic resistance genes in natural water and drinking water. Obviously, there are few studies and reports on this aspect. Although SO4 has strong oxidation ability, it costs a lot to treat high-concentration organic wastewater. Hence, it is necessary to combine SO4●−-based AOPs with traditional biological and physicochemical treatment to achieve the lowest operating cost. It is urgent to carry out research works in this field.
(6) The effectiveness of AOPs depends to a large extent on the generation rate of ROS and the contact with pollutant molecules. It should be paid to the reduction in oxidant dosage and the efficient recycling of catalysts. There are significantly fewer studies based on PS activation than on PMS activation, and new technologies for PS activation are still limited. In fact, commercial PS is much cheaper than PMS, and PS activation produces fewer SO42− ions than PMS activation because PMS only makes up a third of Oxone. Therefore, more attention should be paid to the activation research of PS, especially the research and development of related catalysts and activation methods is very important for the future commercial application.

Author Contributions

Conceptualization, Z.L. and M.Y.; methodology, D.L.; software, Y.S.; validation, F.C., H.L. and Y.D.; formal analysis, Z.L.; investigation, Y.S.; resources, M.Y.; data curation, D.L.; writing—original draft preparation, Z.L.; writing—review and editing, Y.S.; visualization, M.Y.; supervision, D.L.; project administration, F.C.; funding acquisition, M.Y. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by [the National Nonprofit Institute Research Grants of TIWTE], grant number [TKS20220201, TKS20220204] and [research project of Shanghai Science and Technology Commission], grant number [19DZ1204303]. And The APC was funded by [the National Nonprofit Institute Research Grants of TIWTE].

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Wang, J.; Tian, Z.; Huo, Y.B.; Yang, M.; Zheng, X.C.; Zhang, Y. Monitoring of 943 organic micropollutants in wastewater from municipal wastewater treatment plants with secondary and advanced treatment processes. J. Environ. Sci. 2018, 67, 309–317. [Google Scholar] [CrossRef]
  2. Tröger, R.; Klöckner, P.; Ahrens, L.; Wiberg, K. Micropollutants in drinking water from source to tap—Method development and application of a multiresidue screening method. Sci. Total Environ. 2018, 627, 1404–1432. [Google Scholar] [CrossRef]
  3. 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]
  4. Xiao, G.; Xu, T.; Faheem, M.; Xi, Y.; Zhou, T.; Moryani, H.; Bao, J.; Du, J. Evolution of Singlet Oxygen by Activating Peroxydisulfate and Peroxymonosulfate: A Review. Int. J. Environ. Res. Public Health 2021, 18, 3344. [Google Scholar] [CrossRef]
  5. Hu, P.; Long, M. Cobalt-catalyzed sulfate radical-based advanced oxidation: A review on heterogeneous catalysts and applications. Appl. Catal. B Environ. 2016, 181, 103–117. [Google Scholar] [CrossRef]
  6. Ahmed, M.M.; Barbati, S.; Doumenq, P.; Chiron, S. Sulfate radical anion oxidation of diclofenac and sulfamethoxazole for water decontamination. Chem. Eng. J. 2012, 197, 440–447. [Google Scholar] [CrossRef]
  7. Anipsitakis, G.P.; Dionysiou, D.D. Degradation of Organic Contaminants in Water with Sulfate Radicals Generated by the Conjunction of Peroxymonosulfate with Cobalt. Environ. Sci. Technol. 2003, 37, 4790–4797. [Google Scholar] [CrossRef]
  8. Duan, X.; Sun, H.; Shao, Z.; Wang, S. Nonradical reactions in environmental remediation processes: Uncertainty and challenges. Appl. Catal. B Environ. 2018, 224, 973–982. [Google Scholar] [CrossRef]
  9. Bokare, A.D.; Choi, W. Singlet-Oxygen Generation in Alkaline Periodate Solution. Environ. Sci. Technol. 2015, 49, 14392–14400. [Google Scholar] [CrossRef]
  10. Li, Z.; Liu, D.; Zhao, Y.; Li, S.; Wei, X.; Meng, F.; Huang, W.; Lei, Z. Singlet oxygen dominated peroxymonosulfate activation by CuO-CeO2 for organic pollutants degradation: Performance and mechanism. Chemosphere 2019, 233, 549–558. [Google Scholar] [CrossRef]
  11. Mirza-Aghayan, M.; Tavana, M.M.; Boukherroub, R. Direct oxidative synthesis of nitrones from aldehydes and primary anilines using graphite oxide and Oxone. Tetrahedron Lett. 2014, 55, 5471–5474. [Google Scholar] [CrossRef]
  12. Anipsitakis, G.P.; Tufano, T.P.; Dionysiou, D.D. Chemical and microbial decontamination of pool water using activated potassium peroxymonosulfate. Water Res. 2008, 42, 2899–2910. [Google Scholar] [CrossRef]
  13. 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]
  14. House, D.A. Kinetics and Mechanism of Oxidations by Peroxydisulfate. Chem. Rev. 1962, 62, 185–203. [Google Scholar] [CrossRef]
  15. Tang, L.; Ma, X.Y.; Wang, Y.; Zhang, S.; Zheng, K.; Wang, X.C.; Lin, Y. Removal of trace organic pollutants (pharmaceuticals and pesticides) and reduction of biological effects from secondary effluent by typical granular activated carbon. Sci. Total Environ. 2020, 749, 141611. [Google Scholar] [CrossRef]
  16. Bu, Z.; Hou, M.; Li, Z.; Dong, Z.; Zeng, L.; Zhang, P.; Wu, G.; Li, X.; Zhang, Y.; Pan, Y. Fe3+/Fe2+ cycle promoted peroxymonosulfate activation with addition of boron for sulfamethazine degradation: Efficiency and the role of boron. Sep. Purif. Technol. 2022, 298, 121596. [Google Scholar] [CrossRef]
  17. Ding, Y.; Zhu, L.; Wang, N.; Tang, H. Sulfate radicals induced degradation of tetrabromobisphenol A with nanoscaled magnetic CuFe2O4 as a heterogeneous catalyst of peroxymonosulfate. Appl. Catal. B Environ. 2012, 129, 153–162. [Google Scholar] [CrossRef]
  18. 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]
  19. Acero, J.L.; Benítez, F.J.; Real, F.J.; Rodríguez, E. Degradation of selected emerging contaminants by UV-activated persulfate: Kinetics and influence of matrix constituents. Sep. Purif. Technol. 2018, 201, 41–50. [Google Scholar] [CrossRef]
  20. Dong, Z.-Y.; Xu, B.; Hu, C.-Y.; Zhang, T.-Y.; Tang, Y.-L.; Pan, Y.; El-Din, M.G.; Xian, Q.-M.; Gao, N.-Y. The application of UV-C laser in persulfate activation for micropollutant removal: Case study with iodinated X-ray contrast medias. Sci. Total Environ. 2021, 779, 146340. [Google Scholar] [CrossRef]
  21. Ghauch, A.; Tuqan, A.M.; Kibbi, N. Ibuprofen removal by heated persulfate in aqueous solution: A kinetics study. Chem. Eng. J. 2012, 197, 483–492. [Google Scholar] [CrossRef]
  22. Ren, T.-L.; Ma, X.-W.; Wu, X.-Q.; Yuan, L.; Lai, Y.-L.; Tong, Z.-H. Degradation of imidazolium ionic liquids in a thermally activated persulfate system. Chem. Eng. J. 2021, 412, 128624. [Google Scholar] [CrossRef]
  23. Li, B.; Li, L.; Lin, K.; Zhang, W.; Lu, S.; Luo, Q. Removal of 1,1,1-trichloroethane from aqueous solution by a sono-activated persulfate process. Ultrason. Sonochem. 2013, 20, 855–863. [Google Scholar] [CrossRef]
  24. Frontistis, Z.; Mantzavinos, D.; Meriç, S. Degradation of antibiotic ampicillin on boron-doped diamond anode using the combined electrochemical oxidation—Sodium persulfate process. J. Environ. Manag. 2018, 223, 878–887. [Google Scholar] [CrossRef] [PubMed]
  25. Ayoub, G.; Ghauch, A. Assessment of bimetallic and trimetallic iron-based systems for persulfate activation: Application to sulfamethoxazole degradation. Chem. Eng. J. 2014, 256, 280–292. [Google Scholar] [CrossRef]
  26. Wang, S.; Wu, J.; Lu, X.; Xu, W.; Gong, Q.; Ding, J.; Dan, B.; Xie, P. Removal of acetaminophen in the Fe2+/persulfate system: Kinetic model and degradation pathways. Chem. Eng. J. 2018, 358, 1091–1100. [Google Scholar] [CrossRef]
  27. Rastogi, A.; Al-Abed, S.R.; Dionysiou, D.D. Sulfate radical-based ferrous–peroxymonosulfate oxidative system for PCBs degradation in aqueous and sediment systems. Appl. Catal. B Environ. 2009, 85, 171–179. [Google Scholar] [CrossRef]
  28. Hafeez, M.A.; Hong, S.J.; Jeon, J.; Lee, J.; Singh, B.K.; Hyatt, N.C.; Walling, S.A.; Heo, J.; Um, W. Co2+/PMS based sulfate-radical treatment for effective mineralization of spent ion exchange resin. Chemosphere 2022, 287, 132351. [Google Scholar] [CrossRef]
  29. Nfodzo, P.; Choi, H. Triclosan decomposition by sulfate radicals: Effects of oxidant and metal doses. Chem. Eng. J. 2011, 174, 629–634. [Google Scholar] [CrossRef]
  30. 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]
  31. Peng, J.; Zhou, H.; Liu, W.; Ao, Z.; Ji, H.; Liu, Y.; Su, S.; Yao, G.; Lai, B. Insights into heterogeneous catalytic activation of peroxymonosulfate by natural chalcopyrite: pH-dependent radical generation, degradation pathway and mechanism. Chem. Eng. J. 2020, 397, 125387. [Google Scholar] [CrossRef]
  32. Gao, Y.; Han, Y.; Liu, B.; Gou, J.; Feng, D.; Cheng, X. CoFe2O4 nanoparticles anchored on waste eggshell for catalytic oxidation of florfenicol via activating peroxymonosulfate. Chin. Chem. Lett. 2022, 33, 3713–3720. [Google Scholar] [CrossRef]
  33. Shen, Y.; de Vidales, M.J.M.; Gorni, G.; Gómez-Herrero, A.; Fernández-Martínez, F.; Dos Santos-García, A.J. Enhanced performance and recyclability for peroxymonosulfate activation via g-C3N4 supported CoFe layer double oxide. Chem. Eng. J. 2022, 444, 136610. [Google Scholar] [CrossRef]
  34. Chen, X.; Chen, J.; Qiao, X.; Wang, D.; Cai, X. Performance of nano-Co3O4/peroxymonosulfate system: Kinetics and mechanism study using Acid Orange 7 as a model compound. Appl. Catal. B Environ. 2008, 80, 116–121. [Google Scholar] [CrossRef]
  35. Qi, F.; Chu, W.; Xu, B. Catalytic degradation of caffeine in aqueous solutions by cobalt-MCM41 activation of peroxymonosulfate. Appl. Catal. B Environ. 2013, 134–135, 324–332. [Google Scholar] [CrossRef]
  36. Ouyang, M.; Li, X.; Xu, Q.; Tao, Z.; Yao, F.; Huang, X.; Wu, Y.; Wang, D.; Yang, Q.; Chen, Z.; et al. Heterogeneous activation of persulfate by Ag doped BiFeO3 composites for tetracycline degradation. J. Colloid Interface Sci. 2020, 566, 33–45. [Google Scholar] [CrossRef]
  37. Deng, J.; Ye, C.; Cai, A.; Huai, L.; Zhou, S.; Dong, F.; Li, X.; Ma, X. S-doping α-Fe2O3 induced efficient electron-hole separation for enhanced persulfate activation toward carbamazepine oxidation: Experimental and DFT study. Chem. Eng. J. 2021, 420, 129863. [Google Scholar] [CrossRef]
  38. Wu, Z.; Wang, Y.; Xiong, Z.; Ao, Z.; Pu, S.; Yao, G.; Lai, B. Core-shell magnetic Fe3O4@Zn/Co-ZIFs to activate peroxymonosulfate for highly efficient degradation of carbamazepine. Appl. Catal. B Environ. 2020, 277, 119136. [Google Scholar] [CrossRef]
  39. Chen, X.-L.; Li, F.; Zhang, M.; Liu, B.; Chen, H.; Wang, H. Highly dispersed and stabilized Co3O4/C anchored on porous biochar for bisphenol A degradation by sulfate radical advanced oxidation process. Sci. Total Environ. 2021, 777, 145794. [Google Scholar] [CrossRef]
  40. Xu, R.; Li, M.; Zhang, Q. Collaborative optimization for the performance of ZnO/biochar composites on persulfate activation through plant enrichment-pyrolysis method. Chem. Eng. J. 2021, 429, 132294. [Google Scholar] [CrossRef]
  41. Ghorbanian, Z.; Asgari, G.; Samadi, M.T.; Seid-Mohammadi, A. Removal of 2,4 dichlorophenol using microwave assisted nanoscale zero-valent copper activated persulfate from aqueous solutions: Mineralization, kinetics, and degradation pathways. J. Mol. Liq. 2019, 296, 111873. [Google Scholar] [CrossRef]
  42. Zou, X.; Zhou, T.; Mao, J.; Wu, X. Synergistic degradation of antibiotic sulfadiazine in a heterogeneous ultrasound-enhanced Fe0/persulfate Fenton-like system. Chem. Eng. J. 2014, 257, 36–44. [Google Scholar] [CrossRef]
  43. Zhang, J.; Shao, X.; Shi, C.; Yang, S. Decolorization of Acid Orange 7 with peroxymonosulfate oxidation catalyzed by granular activated carbon. Chem. Eng. J. 2013, 232, 259–265. [Google Scholar] [CrossRef]
  44. Pervez, N.; He, W.; Zarra, T.; Naddeo, V.; Zhao, Y. New Sustainable Approach for the Production of Fe3O4/Graphene Oxide-Activated Persulfate System for Dye Removal in Real Wastewater. Water 2020, 12, 733. [Google Scholar] [CrossRef]
  45. Hong, Y.; Luo, Z.; Zhang, N.; Qu, L.; Zheng, M.; Suara, M.A.; Chelme-Ayala, P.; Zhou, X.; El-Din, M.G. Decomplexation of Cu(II)-EDTA by synergistic activation of persulfate with alkali and CuO: Kinetics and activation mechanism. Sci. Total Environ. 2022, 817, 152793. [Google Scholar] [CrossRef] [PubMed]
  46. Zhang, B.-T.; Zhang, Y.; Teng, Y.; Fan, M. Sulfate Radical and Its Application in Decontamination Technologies. Crit. Rev. Environ. Sci. Technol. 2014, 45, 1756–1800. [Google Scholar] [CrossRef]
  47. Liang, C.; Bruell, C.J. Thermally Activated Persulfate Oxidation of Trichloroethylene: Experimental Investigation of Reaction Orders. Ind. Eng. Chem. Res. 2008, 47, 2912–2918. [Google Scholar] [CrossRef]
  48. Liang, C.J.; Bruell, C.J.; Marley, M.C.; Sperry, K.L. Thermally Activated Persulfate Oxidation of Trichloroethylene (TCE) and 1,1,1-Trichloroethane (TCA) in Aqueous Systems and Soil Slurries. Soil Sediment Contam. Int. J. 2003, 12, 207–228. [Google Scholar] [CrossRef]
  49. Yang, S.; Wang, P.; Yang, X.; Shan, L.; Zhang, W.; Shao, X.; Niu, R. Degradation efficiencies of azo dye Acid Orange 7 by the interaction of heat, UV and anions with common oxidants: Persulfate, peroxymonosulfate and hydrogen peroxide. J. Hazard. Mater. 2010, 179, 552–558. [Google Scholar] [CrossRef]
  50. He, X.; de la Cruz, A.A.; Dionysiou, D.D. Destruction of cyanobacterial toxin cylindrospermopsin by hydroxyl radicals and sulfate radicals using UV-254nm activation of hydrogen peroxide, persulfate and peroxymonosulfate. J. Photochem. Photobiol. A Chem. 2013, 251, 160–166. [Google Scholar] [CrossRef]
  51. George, C.; Chovelon, J.-M. A laser flash photolysis study of the decay of SO4− and Cl2− radical anions in the presence of Cl− in aqueous solutions. Chemosphere 2002, 47, 385–393. [Google Scholar] [CrossRef]
  52. Chen, X.; Qiao, X.; Wang, D.; Lin, J.; Chen, J. Kinetics of oxidative decolorization and mineralization of Acid Orange 7 by dark and photoassisted Co2+-catalyzed peroxymonosulfate system. Chemosphere 2006, 67, 802–808. [Google Scholar] [CrossRef] [PubMed]
  53. Barceloux, D.G. Cobalt. J. Toxicol. Clin. Toxicol. 1999, 37, 201–216. [Google Scholar] [CrossRef]
  54. Rodriguez, S.; Vasquez, L.; Costa, D.; Romero, A.; Santos, A. Oxidation of Orange G by persulfate activated by Fe(II), Fe(III) and zero valent iron (ZVI). Chemosphere 2014, 101, 86–92. [Google Scholar] [CrossRef]
  55. Li, C.-X.; Wang, Y.-J.; Chen, C.-B.; Fu, X.-Z.; Cui, S.; Lu, J.-Y.; Liu, H.-Q.; Li, W.-W. Interactions between chlorophenols and peroxymonosulfate: pH dependency and reaction pathways. Sci. Total Environ. 2019, 664, 133–139. [Google Scholar] [CrossRef] [PubMed]
  56. 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]
  57. Duan, X.; Sun, H.; Kang, J.; Wang, Y.; Indrawirawan, S.; Wang, S. Insights into Heterogeneous Catalysis of Persulfate Activation on Dimensional-Structured Nanocarbons. ACS Catal. 2015, 5, 4629–4636. [Google Scholar] [CrossRef]
  58. Jans, U.; Hoigné, J. Activated Carbon and Carbon Black Catalyzed Transformation of Aqueous Ozone into OH-Radicals. Ozone: Sci. Eng. 1998, 20, 67–90. [Google Scholar] [CrossRef]
  59. Zhou, Y.; Jiang, J.; Gao, Y.; Pang, S.-Y.; Yang, Y.; Ma, J.; Gu, J.; Li, J.; Wang, Z.; Wang, L.-H.; et al. Activation of peroxymonosulfate by phenols: Important role of quinone intermediates and involvement of singlet oxygen. Water Res. 2017, 125, 209–218. [Google Scholar] [CrossRef]
  60. Schweitzer, C.; Schmidt, R. Physical Mechanisms of Generation and Deactivation of Singlet Oxygen. Chem. Rev. 2003, 103, 1685–1758. [Google Scholar] [CrossRef]
  61. 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] [PubMed]
  62. Krieger-Liszkay, A. Singlet oxygen production in photosynthesis. J. Exp. Bot. 2005, 56, 337–346. [Google Scholar] [CrossRef] [PubMed]
  63. Hotze, E.M.; Badireddy, A.R.; Chellam, S.; Wiesner, M.R. Mechanisms of Bacteriophage Inactivation via Singlet Oxygen Generation in UV Illuminated Fullerol Suspensions. Environ. Sci. Technol. 2009, 43, 6639–6645. [Google Scholar] [CrossRef] [PubMed]
  64. Yuan, Y.; Zhang, C.-J.; Xu, S.; Liu, B. A self-reporting AIE probe with a built-in singlet oxygen sensor for targeted photodynamic ablation of cancer cells. Chem. Sci. 2015, 7, 1862–1866. [Google Scholar] [CrossRef]
  65. 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. 2018, 53, 307–315. [Google Scholar] [CrossRef]
  66. Rkanofsky, J. Assay for singlet-oxygen generation by peroxidases using 1270-nm chemiluminescence. Methods Enzymol. 2000, 319, 59–67. [Google Scholar]
  67. Qu, X.; Alvarez, P.J.J.; Li, Q. Photochemical Transformation of Carboxylated Multiwalled Carbon Nanotubes: Role of Reactive Oxygen Species. Environ. Sci. Technol. 2013, 47, 14080–14088. [Google Scholar] [CrossRef]
  68. Bartusik, D.; Aebisher, D.; Lyons, A.M.; Greer, A. Bacterial Inactivation by a Singlet Oxygen Bubbler: Identifying Factors Controlling the Toxicity of 1O2 Bubbles. Environ. Sci. Technol. 2012, 46, 12098–12104. [Google Scholar] [CrossRef]
  69. Romero, O.C.; Straub, A.P.; Kohn, T.; Nguyen, T.H. Role of Temperature and Suwannee River Natural Organic Matter on Inactivation Kinetics of Rotavirus and Bacteriophage MS2 by Solar Irradiation. Environ. Sci. Technol. 2011, 45, 10385–10393. [Google Scholar] [CrossRef]
  70. Mostafa, S.; Rosario-Ortiz, F.L. Singlet Oxygen Formation from Wastewater Organic Matter. Environ. Sci. Technol. 2013, 47, 8179–8186. [Google Scholar] [CrossRef]
  71. Guo, X.; Li, Q.; Zhang, M.; Long, M.; Kong, L.; Zhou, Q.; Shao, H.; Hu, W.; Wei, T. Enhanced photocatalytic performance of N-nitrosodimethylamine on TiO2 nanotube based on the role of singlet oxygen. Chemosphere 2015, 120, 521–526. [Google Scholar] [CrossRef] [PubMed]
  72. Zhang, T.; Ding, Y.; Tang, H. Generation of singlet oxygen over Bi(V)/Bi(III) composite and its use for oxidative degradation of organic pollutants. Chem. Eng. J. 2015, 264, 681–689. [Google Scholar] [CrossRef]
  73. Yu, K.; Yang, S.; Boyd, S.A.; Chen, H.; Sun, C. Efficient degradation of organic dyes by BiAgxOy. J. Hazard. Mater. 2011, 197, 88–96. [Google Scholar] [CrossRef] [PubMed]
  74. Hao, Y.-J.; Liu, B.; Tian, L.-G.; Li, F.-T.; Ren, J.; Liu, S.-J.; Liu, Y.; Zhao, J.; Wang, X.-J. Synthesis of {111} Facet-Exposed MgO with Surface Oxygen Vacancies for Reactive Oxygen Species Generation in the Dark. ACS Appl. Mater. Interfaces 2017, 9, 12687–12693. [Google Scholar] [CrossRef]
  75. Stephenson, L.M.; McClure, D.E. Mechanisms in phosphite ozonide decomposition to phosphate esters and singlet oxygen. J. Am. Chem. Soc. 1973, 95, 3074–3076. [Google Scholar] [CrossRef]
  76. Foote, C.S.; Wexler, S.; Ando, W.; Higgins, R. Chemistry of singlet oxygen. IV. Oxygenations with hypochlorite-hydrogen peroxide. J. Am. Chem. Soc. 1968, 90, 975–981. [Google Scholar] [CrossRef]
  77. Du, J.; Xiao, G.; Xi, Y.; Zhu, X.; Su, F.; Kim, S.H. Periodate activation with manganese oxides for sulfanilamide degradation. Water Res. 2020, 169, 115278. [Google Scholar] [CrossRef]
  78. 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]
  79. Yang, F.; Huang, Y.; Fang, C.; Xue, Y.; Ai, L.; Liu, J.; Wang, Z. Peroxymonosulfate/base process in saline wastewater treatment: The fight between alkalinity and chloride ions. Chemosphere 2018, 199, 84–88. [Google Scholar] [CrossRef]
  80. Lou, X.; Fang, C.; Geng, Z.; Jin, Y.; Xiao, D.; Wang, Z.; Liu, J.; Guo, Y. Significantly enhanced base activation of peroxymonosulfate by polyphosphates: Kinetics and mechanism. Chemosphere 2017, 173, 529–534. [Google Scholar] [CrossRef]
  81. Peng, G.; You, W.; Zhou, W.; Zhou, G.; Qi, C.; Hu, Y. Activation of peroxymonosulfate by phosphite: Kinetics and mechanism for the removal of organic pollutants. Chemosphere 2020, 266, 129016. [Google Scholar] [CrossRef] [PubMed]
  82. Rao, L.; Yang, Y.; Chen, L.; Liu, X.; Chen, H.; Yao, Y.; Wang, W. Highly efficient removal of organic pollutants via a green catalytic oxidation system based on sodium metaborate and peroxymonosulfate. Chemosphere 2019, 238, 124687. [Google Scholar] [CrossRef] [PubMed]
  83. Nie, M.; Zhang, W.; Yan, C.; Xu, W.; Wu, L.; Ye, Y.; Hu, Y.; Dong, W. Enhanced removal of organic contaminants in water by the combination of peroxymonosulfate and carbonate. Sci. Total Environ. 2018, 647, 734–743. [Google Scholar] [CrossRef] [PubMed]
  84. Othman, I.; Zain, J.H.; Abu Haija, M.; Banat, F. Catalytic activation of peroxymonosulfate using CeVO4 for phenol degradation: An insight into the reaction pathway. Appl. Catal. B Environ. 2020, 266, 118601. [Google Scholar] [CrossRef]
  85. Chi, Y.; Wang, P.; Lin, M.; Lin, C.; Gao, M.; Zhao, C.; Wu, X. Manganese oxides activated peroxymonosulfate for ciprofloxacin removal: Effect of oxygen vacancies and chemical states. Chemosphere 2022, 299, 134437. [Google Scholar] [CrossRef]
  86. Jawad, A.; Zhan, K.; Wang, H.; Shahzad, A.; Zeng, Z.; Wang, J.; Zhou, X.; Ullah, H.; Chen, Z.; Chen, Z. Tuning of Persulfate Activation from a Free Radical to a Nonradical Pathway through the Incorporation of Non-Redox Magnesium Oxide. Environ. Sci. Technol. 2020, 54, 2476–2488. [Google Scholar] [CrossRef]
  87. Li, Z.; Sun, Y.; Huang, W.; Xue, C.; Zhu, Y.; Wang, Q.; Liu, D. Innovatively employing magnetic CuO nanosheet to activate peroxymonosulfate for the treatment of high-salinity organic wastewater. J. Environ. Sci. 2020, 88, 46–58. [Google Scholar] [CrossRef]
  88. Yu, L.; Zhang, G.; Liu, C.; Lan, H.; Liu, H.; Qu, J. Interface Stabilization of Undercoordinated Iron Centers on Manganese Oxides for Nature-Inspired Peroxide Activation. ACS Catal. 2018, 8, 1090–1096. [Google Scholar] [CrossRef]
  89. Yu, R.; Zhao, J.; Zhao, Z.; Cui, F. Copper substituted zinc ferrite with abundant oxygen vacancies for enhanced ciprofloxacin degradation via peroxymonosulfate activation. J. Hazard. Mater. 2019, 390, 121998. [Google Scholar] [CrossRef]
  90. Shao, P.; Tian, J.; Yang, F.; Duan, X.; Gao, S.; Shi, W.; Luo, X.; Cui, F.; Luo, S.; Wang, S. Identification and Regulation of Active Sites on Nanodiamonds: Establishing a Highly Efficient Catalytic System for Oxidation of Organic Contaminants. Adv. Funct. Mater. 2018, 28, 1705295. [Google Scholar] [CrossRef]
  91. Gao, Y.; Chen, Z.; Zhu, Y.; Li, T.; Hu, C. New Insights into the Generation of Singlet Oxygen in the Metal-Free Peroxymonosulfate Activation Process: Important Role of Electron-Deficient Carbon Atoms. Environ. Sci. Technol. 2019, 54, 1232–1241. [Google Scholar] [CrossRef] [PubMed]
  92. Gao, Y.; Li, T.; Zhu, Y.; Chen, Z.; Liang, J.; Zeng, Q.; Lyu, L.; Hu, C. Highly nitrogen-doped porous carbon transformed from graphitic carbon nitride for efficient metal-free catalysis. J. Hazard. Mater. 2019, 393, 121280. [Google Scholar] [CrossRef] [PubMed]
  93. Ma, W.; Wang, N.; Fan, Y.; Tong, T.; Han, X.; Du, Y. Non-radical-dominated catalytic degradation of bisphenol A by ZIF-67 derived nitrogen-doped carbon nanotubes frameworks in the presence of peroxymonosulfate. Chem. Eng. J. 2018, 336, 721–731. [Google Scholar] [CrossRef]
  94. Li, Z.; Liu, D.; Huang, W.; Wei, X.; Huang, W. Biochar supported CuO composites used as an efficient peroxymonosulfate activator for highly saline organic wastewater treatment. Sci. Total Environ. 2020, 721, 137764. [Google Scholar] [CrossRef]
  95. Liang, P.; Zhang, C.; Duan, X.; Sun, H.; Liu, S.; Tade, M.O.; Wang, S. An insight into metal organic framework derived N-doped graphene for the oxidative degradation of persistent contaminants: Formation mechanism and generation of singlet oxygen from peroxymonosulfate. Environ. Sci. Nano 2017, 4, 315–324. [Google Scholar] [CrossRef]
  96. Montgomery, R.E. Catalysis of peroxymonosulfate reactions by ketones. J. Am. Chem. Soc. 1974, 96, 7820–7821. [Google Scholar] [CrossRef]
  97. Lange, A.; Brauer, H.-D. On the formation of dioxiranes and of singlet oxygen by the ketone-catalysed decomposition of Caro’s acid. J. Chem. Soc. Perkin Trans. 2 1996, 5, 805–811. [Google Scholar] [CrossRef]
  98. 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]
  99. Liu, Y.; Guo, H.; Zhang, Y.; Tang, W.; Cheng, X.; Li, W. Heterogeneous activation of peroxymonosulfate by sillenite Bi25FeO40: Singlet oxygen generation and degradation for aquatic levofloxacin. Chem. Eng. J. 2018, 343, 128–137. [Google Scholar] [CrossRef]
  100. Yang, S.; Wu, P.; Liu, J.; Chen, M.; Ahmed, Z.; Zhu, N. Efficient removal of bisphenol A by superoxide radical and singlet oxygen generated from peroxymonosulfate activated with Fe0-montmorillonite. Chem. Eng. J. 2018, 350, 484–495. [Google Scholar] [CrossRef]
  101. Liu, L.; Li, Y.; Li, W.; Zhong, R.; Lan, Y.; Guo, J. The efficient degradation of sulfisoxazole by singlet oxygen (1O2) derived from activated peroxymonosulfate (PMS) with Co3O4–SnO2/RSBC. Environ. Res. 2020, 187, 109665. [Google Scholar] [CrossRef] [PubMed]
  102. Zhang, Q. Mechanism and Performance of Singlet Oxygen Dominated Peroxymonosulfate Activation on CoOOH Nanopar-ticles for 2,4-Dichlorophenol Degradation in Water. J. Hazard. Mater. 2020, 8. [Google Scholar]
  103. 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]
  104. Yu, J.; Feng, H.; Tang, L.; Pang, Y.; Zeng, G.; Lu, Y.; Dong, H.; Wang, J.; Liu, Y.; Feng, C.; et al. Metal-free carbon materials for persulfate-based advanced oxidation process: Microstructure, property and tailoring. Prog. Mater. Sci. 2020, 111, 100654. [Google Scholar] [CrossRef]
  105. Ren, W.; Xiong, L.; Nie, G.; Zhang, H.; Duan, X.; Wang, S. Insights into the Electron-Transfer Regime of Peroxydisulfate Activation on Carbon Nanotubes: The Role of Oxygen Functional Groups. Environ. Sci. Technol. 2019, 54, 1267–1275. [Google Scholar] [CrossRef]
  106. Duan, X.; Ao, Z.; Zhou, L.; Sun, H.; Wang, G.; Wang, S. Occurrence of radical and nonradical pathways from carbocatalysts for aqueous and nonaqueous catalytic oxidation. Appl. Catal. B Environ. 2016, 188, 98–105. [Google Scholar] [CrossRef]
  107. Chen, X.; Oh, W.-D.; Lim, T.-T. Graphene- and CNTs-based carbocatalysts in persulfates activation: Material design and catalytic mechanisms. Chem. Eng. J. 2018, 354, 941–976. [Google Scholar] [CrossRef]
  108. Gao, Y.; Zhu, Y.; Lyu, L.; Zeng, Q.; Xing, X.; Hu, C. Electronic Structure Modulation of Graphitic Carbon Nitride by Oxygen Doping for Enhanced Catalytic Degradation of Organic Pollutants through Peroxymonosulfate Activation. Environ. Sci. Technol. 2018, 52, 14371–14380. [Google Scholar] [CrossRef]
  109. Huang, B.-C.; Jiang, J.; Huang, G.-X.; Yu, H.-Q. Sludge biochar-based catalysts for improved pollutant degradation by activating peroxymonosulfate. J. Mater. Chem. A 2018, 6, 8978–8985. [Google Scholar] [CrossRef]
Figure 1. The activation reactions of PMS (a) and PS (b) by energy.
Figure 1. The activation reactions of PMS (a) and PS (b) by energy.
Catalysts 12 01092 g001
Figure 2. Mechanism diagram of Co catalyst activated PMS for degradation of organic pollutants.
Figure 2. Mechanism diagram of Co catalyst activated PMS for degradation of organic pollutants.
Catalysts 12 01092 g002
Figure 3. The mechanism diagram of PMS and PS activation by metal ions and metal oxides.
Figure 3. The mechanism diagram of PMS and PS activation by metal ions and metal oxides.
Catalysts 12 01092 g003
Figure 4. Schematic diagram of organic pollutants degradation by non-radical AOPs (Oxidant can be PMS, PS, H2O2 or O3).
Figure 4. Schematic diagram of organic pollutants degradation by non-radical AOPs (Oxidant can be PMS, PS, H2O2 or O3).
Catalysts 12 01092 g004
Table 1. Properties of PMS and PS.
Table 1. Properties of PMS and PS.
PropertiesPS (Take Potassium Persulfate as an Example)PMS (Take Potassium Peroxymonosulfate as an Example)
CAS Registry Number7727-21-110058-23-8
Chemical formulaK2S2O8KHSO5
Molecular mass270.309614.738
Solubility in water (20 °C)520 g/L>250 g/L
Redox potential2.01 V1.82 V
Table 2. PMS and PS radical activation with various method for the removal of pollutants.
Table 2. PMS and PS radical activation with various method for the removal of pollutants.
Reaction SystemPollutantConditionsReactivityDominant ROSRef.
UV(254 nm)/PMSbenzoic acid (BA)[BA] = 9.90 μM; [PMS] = 100 μM as 1/2 Oxone;
pH = 11.
>90% with 10 minSO4●− and OH[18]
UV(254 nm)/PSChlorophene[Chlorophene] = 1 μM, [PS] = 50 μM, pH = 7, UV intensity (254 nm,15 W) = 4.23 mWcm−2. 100% with 5 minSO4●−[19]
UV-C laser/PSIohexol (IOX)IUV = 25 mW/cm2, [PS] = 1.0 mM, [IOX] = 10 μM, initial pH = 7.0 ± 0.5, temperature 25 °C.93.8% within only 40 sSO4●−[20]
Heat/PS Ibuprofen (IBU)[phosphate buffer] = 0.09 M, pH = 7.0, [IBU] = 20.36 μM, [PS] = 1.0 mM, T = 70 °C.100% with 20 minSO4●− and OH[21]
Heat/PS1-alkyl-3-methylimidazolium bromides (C4mimBr)[C4mimBr] = 0.1 mM, [PS] = 10 mM, T = 60 °C, pH = 7, V = 100 mL.100% with 120 minSO4●−[22]
Ultrasound/PS1,1,1-trichloroethane (TCA)[TCA] = 25.0 mg/L, [PS] = 250.0 mg/L pH 7.0, T = 20 ± 2 °C, ultrasound: 400 kHz, 100 W.100% with 120 minSO4●−[23]
Electrolysis(boron-doped diamond anode)/PSAmpicillin[Ampicillin] = 1.1 mg/L, [PS] = 250 mg/L, current density (BDD anode) = 25 mAcm−2.100% with 120 minSO4●− and OH[24]
Fe0/PSsulfamethoxazole (SMX)[SMX] = 39.5 μM, [PS] = 1.0 mM, [Fe0] = 2.23 mM; m (Fe0) = 2.5 mg, pH = 3.52. 100% with 30 minSO4●−[25]
Fe2+/PSAcetaminophen (ACT)[ACT] = 0.05 mM, [Fe2+] = 1 mM, [PS] = 0.8 mM, pH = 3, T = 20 °C.70% with 30 minSO4●−[26]
Fe(II)/PMS2-chlorobiphenyl (2-CB)[2-CB] = 0.0212 mM; [Fe(II)] = 0.11 mM; [PMS]0 = 0.11 mM. 90% with 240 minSO4●−[27]
Co2+/PMSnuclear grade cationic IRN-77 resininitial pH = 9, [Co2+] = 4 mM, [PMS] = 60 mM T = 60 °C.∼90% COD removal (1000 mg/L) with 60 minSO4●− and OH[28]
Cu2+/PMSTriclosan (TCS)[TCS] = 9 mg/L (0.031 mM), initial pH = 7, molar ratio of oxidant to metal = 1:1, molar ratio of oxidant to triclosan = 5:1. 95% with 10 minSO4●−[29]
Ru3+/PMS2,4-dichlorophenol (2,4-DCP)[2,4-DCP] = 0.311 mM, [RuCl3·xH2O] = 2.553 mM, [KHSO5] = 1.244 mM, pH = 7.98% in less than 1 minSO4●−[30]
Natural chalcopyrite/PMSBisphenol SBisphenol S = 25 μM, chalcopyrite = 2 g/L, PMS = 0.4 mM, initial pH = 6.2, T = 303 K.83% with 30 minSO4•− and OH[31]
Eggshell-loaded CoFe2O4./PMSFlorfenicol (FF)CoFe2O4/eggshell = 0.4 g/L, [PMS] = 0.96 mmol/L, [FF] = 10 mg/L, T = 30 °C, initial pH = 6.61.96.8% within 40 min SO4•– and OH[32]
CoFe layered double oxide/g-C3N4/PMSparacetamol[catalyst] = 0.2 g/L, [PMS] = 0.5 mM, [paracetamol] = 10 mg/L, temperature = 25 ± 0.5 °C, initial pH = 7 ± 0.2.100% in less than 10 min SO4●−[33]
Co3O4/PMSAcid Orange 7[AO7] = 0.2 mM, [PMS] = 2 mM and [nano-Co3O4] = 0.5 g/L, pH = 7.100% with 60 minSO4●−[34]
Co-MCM41Caffeine (CAF)[CAF] = 0.05 mM, [PMS] = 0.2 mM, [catalyst] = 200 mg/L, pH = 7.10.100% with 15 minSO4•– and OH[35]
Ag0.4-BiFeO3/PStetracycline (TC)[catalysts] = 300 mg/L, [PS] = 5 mM, [TC] = 10 mg/L, pH = 4.5, T = 298 K.91% with 60 minSO4●− and OH[36]
S-doped α-Fe2O3/PScarbamazepine (CBZ)[CBZ] = 2 mg/L, [PS]0 = 0.2 mM, [catalyst] = 0.2 g/L, T = 25 ± 1 °C and initial pH = 6.8 ± 0.5.93.13% with 30 minSO4●−[37]
Fe3O4@Zn/Co-ZIFs/PMScarbamazepine (CBZ)[CBZ] = 5 mg/L, [catalyst] = 25 mg/L, [PMS] = 0.4 mM, initial pH = 6.8, T = 30 °C.100% with 30 minSO4●−[38]
Co3O4/C-BC/PMSBisphenol A (BPA)[catalyst] = 0.3 g/L, [pollutant] = 20 mg/L, [PMS] = 1.0 mmol/L, [pH] = 7.0, [T] = 30 °C.100% within less than 30 minSO4●−[39]
ZnO/biochar/PStetracycline hydrochloride (TC)[TC] = 0.05 g/L, [ZnO200/BC] = 0.1 g/L, [PS] = 1.0 mM, pH = 7.0 ± 0.1, T = 25 ± 2 °C.44.98% with 50 minSO4●− and OH[40]
microwave irradiation(MW)/CuO/PS2,4-dichlorophenol (2,4-DCP)[PS] = 0.4 g/L, [CuO] = 40 mg/L, (2,4-DCP) = 50 mg/L, initial pH = 9, MW power intensity = 180 W.>98% with 90 minSO4●−[41]
Ultrasound/Fe0/PSSulfadiazine (SD)[SD] = 20 mg/L, [Fe0] = 0.92 Mm, [PS] = 1.84 mM, US input power = 90 W, initial pH = 7, room temperature.99.1% with 60 minSO4●−[42]
granular activated carbon (GAC)/PMSAcid Orange 7 (AO7)[AO7] = 20 mg/L, [PMS]: [AO7] = 100:1, [GAC] = 1.0 g/L, without pH adjustment, T = 20 ± 0.5 °C.85% with 5 hSO4●−[43]
Fe3O4@Graphene oxide (GO)/PSRhodamine B (RhB)[RhB] = 20 ppm, [Fe3O4@GO] = 500 mg/L, [PS] = 1.5 mM, pH = 4.34, T = 20 °C.89% with 120 minSO4●−[44]
alkali and CuO/PSCu-ethylenediamine tetraacetic acid (Cu(II)-EDTA)[Cu(II)-EDTA] = 3.14 mM, [PS]/[Cu(II)-EDTA] = 15:1, [CuO] = 2 g/L, pH maintained at 11.Nearly 100% with 120 minSO4•−, OH and O2•−[45]
Table 3. The removal of pollutants via 1O2 oxidation generated by PMS and PS activation.
Table 3. The removal of pollutants via 1O2 oxidation generated by PMS and PS activation.
Reaction SystemPollutantConditionsReactivityDominant ROSRef.
Chlorophenols/PMS2, 4, 6-trichlorophenol (2, 4, 6-TCP)[2, 4, 6-TCP] = 25 μM, [PMS] = 0.325 mM, T = 20 ± 2 °C, pH = 9.95% with 60 min1O2[55]
benzoquinone (BQ)/PMSsulfamethoxazole (SMX)[PMS] = 0.44 mM,
[SMX] = 8 μM, [BQ] = 10 μM, pH = 10, T = 25 °C.
86% with 3 min1O2[78]
PMS/baseMethylene blue (MB)[PMS] = 1 mM, [MB] = 0.03 mM, [NaOH] = 2 mM, without pH adjustment100% with 120 min1O2[79]
pyrophosphate (PA)/PMSAcid Orange 7 (AO7)[AO7] = 50 μΜ, [PA] = 0.1 M, [PMS] = 2.5 mM, pH 9.5.Nearly 100% with 10 min1O2[80]
Phosphite (HPO32−)/PMSAcid Orange 7 (AO7)[AO7] = 20 mg/L, [PMS] = 10.0 mM, [HPO32−] = 25.0 mM, pH = 7.06.82.1% within 60 min1O2[81]
NaBO2/PMSAcid Red 1[AR1] = 50 μM, [PMS] = 3 mM, [NaBO2] = 7.5 mM, initial pH = 7.0, T = 30 °C.97.8% within 10 min1O2[82]
carbonate (CO32−)/PMSAcid Orange 7 (AO7)[AO7] = 0.05 mM, [CO32−] = 5 mM, [PMS] = 1 mM.100% within 40 minO2 and 1O2[83]
β-MnO2/PSphenol[β-MnO2] = 400 mg/L, [PS] = 4 mM, [phenol] = 100 μM, pH buffered around 6.5.Over 99% with 180 min1O2[65]
CeVO4/PMSphenolPhenol = 100 ppm, CeVO4 = 1.0 g/L, PMS = 2.0 g/L.100% with 80 min1O2[84]
ε-MnO2/PMSCiprofloxacin (CIP)[CIP] = 10 ppm, [catalyst] = 0.1 g/L, [PMS] = 1 mM, without pH adjustment.84.3% with 120 min1O2[85]
CuOMgO/Fe3O4/PMPMS4-chlorophenol (4-CP)[4-CP] = 40 ppm, [PMS] = 2 mM, [catalyst] = 0.2 g/L, T = 30 °C.100% within 40 min1O2[86]
CuO-CeO2/PMSRhodamine B (RhB)[PMS] = 1.6 mM, [catalyst] = 0.4 g/L, initial pH 7, [RhB] = 0.1 mM.100% within 60 min1O2[10]
CuO-Fe3O4/PMSAcid Orange 7 (AO7)[Na2SO4] = 0.2 M, [PMS] = 2 mM,
[catalyst] = 0.1 g/L, initial pH = 7, [AO7] = 0.2 mM.
95.81% within 30 min1O2[87]
Iron centers on manganese oxides (FeMn-350)/PMSbisphenol A (BPA)[BPA] = 80 mg/L, [catalyst] = 0.5 g/L, [Oxone] = 0.6 mM, pH = 7.5.Nearly 100% within 25 min1O2[88]
copper substituted zinc ferrite (ZCFO)/PMSCiprofloxacin (CIP)[ZCFO] = 0.1 g/L, [PMS] = 2.5 mM, [CIP] = 10 mg/L and pH = 7.5.96.6% with 15 minO2 and 1O2[89]
carbon nanotubes (CNTs)/PS2,4-dichlorophenol (2,4-DCP)[2,4-DCP] = 0.031 mM, PS/2,4-DCP molar ratio = 1/1, [CNTs] = 0.10 g/L, pH = 6.50 ± 0.05.95.9% within 30 min1O2[61]
Nanodiamonds(AND/800)/PMS4-chlorophenol (4-CP)[4-CP] = 0.16 mM, [PMS] = 0.25 mM, [AND/800] = 0.1 g/L, T = 20 ± 2 °C.81% with 30 min1O2[90]
nitrogen-doped carbon nanosheets (NCN-900)/PMSBisphenol A (BPA)[BPA] = 0.1 mM, [NCN-900] = 0.1 g/L, [PMS] = 2 mM, T = 30 °C, initial pH = 6.7.100% within 2 min1O2[91]
N-doped porous carbon (NC1.0)/PMSBisphenol A (BPA)[BPA] = 0.1 mM, [(NC1.0] = 0.2 g/L, [PMS] = 2 mM, T =30 °C, initial pH = 6.7.100% within 15 min1O2[92]
nitrogen-doped carbon nanotubes frameworks (NCNTFs)/PMSBisphenol A (BPA)[BPA] = 25 mg/L, [NCNTFs-800] = 0.05 g/L, [Oxone] = 0.40 g/L, temperature = 20 °C.97.3% with 30 min1O2[93]
CuO-Biochar/PMSAtrazine (ATZ)][Na2SO4] = 200 mM, [PMS] = 2 mM, [catalyst] = 0.2 g/L, [ATZ] = 0.1 mM, initial
pH = 7.
100% within 30 min1O2[94]
N-doped graphene/PMSphenol[catalyst] = 100 mg/L, [PMS] = 3.25 mM, [phenol] = 50 ppm, T = 25 °C.100% within 30 min1O2[95]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Li, Z.; Sun, Y.; Liu, D.; Yi, M.; Chang, F.; Li, H.; Du, Y. A Review of Sulfate Radical-Based and Singlet Oxygen-Based Advanced Oxidation Technologies: Recent Advances and Prospects. Catalysts 2022, 12, 1092. https://doi.org/10.3390/catal12101092

AMA Style

Li Z, Sun Y, Liu D, Yi M, Chang F, Li H, Du Y. A Review of Sulfate Radical-Based and Singlet Oxygen-Based Advanced Oxidation Technologies: Recent Advances and Prospects. Catalysts. 2022; 12(10):1092. https://doi.org/10.3390/catal12101092

Chicago/Turabian Style

Li, Zhendong, Yanmei Sun, Dongfang Liu, Malan Yi, Fang Chang, Huiting Li, and Yunyi Du. 2022. "A Review of Sulfate Radical-Based and Singlet Oxygen-Based Advanced Oxidation Technologies: Recent Advances and Prospects" Catalysts 12, no. 10: 1092. https://doi.org/10.3390/catal12101092

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