Next Article in Journal
Improvement of Water Quality by Light-Emitting Diode Illumination at the Bottom of a Field Experimental Pond
Next Article in Special Issue
A Review of the Treatment Process of Perfluorooctane Compounds in the Waters: Adsorption, Flocculation, and Advanced Oxidative Process
Previous Article in Journal
Simulating Dynamics and Ecology of the Sea Ice of the White Sea by the Coupled Ice–Ocean Numerical Model
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Review of Advanced Oxidation Processes Based on Peracetic Acid for Organic Pollutants

1
School of Environment and Architecture, University of Shanghai for Science and Technology, Shanghai 200093, China
2
Université Nice Sophia Antipolis, Champagne Saint-Dizier, CEDEX 02, 06108 Nice, France
*
Author to whom correspondence should be addressed.
Water 2022, 14(15), 2309; https://doi.org/10.3390/w14152309
Submission received: 30 June 2022 / Revised: 18 July 2022 / Accepted: 21 July 2022 / Published: 25 July 2022

Abstract

:
In recent years, the removal of organic pollutants from water and wastewater has attracted more attention to different advanced oxidation processes (AOPs). There has been increasing interest in using peroxyacetic acid (PAA), an emerging oxidant with low or no toxic by-products, yet the promotion and application are limited by unclear activation mechanisms and complex preparation processes. This paper synthesized the related research results reported on the removal of organic pollutants by PAA-based AOPs. Based on the research of others, this paper not only introduced the preparation method and characteristics of PAA but also summarized the mechanism and reactivity of PAA activated by the free radical pathway and discussed the main influencing factors. Furthermore, the principle and application of the newly discovered methods of non-radical activation of PAA in recent years were also reviewed for the first time. Finally, the shortcomings and development of PAA-based AOPs were discussed and prospected. This review provides a reference for the development of activated PAA technology that can be practically applied to the treatment of organic pollutants in water.

1. Introduction

In recent decades, water shortage and pollution have drawn worldwide attention with the development of science and technology, the sharp increase in population and the acceleration of urbanization [1]. It was reported that some refractory organics in found water, such as dyes, leather, pharmaceuticals, rubber and pesticides, come from chemical industries. These materials are synthetic organic products with stable chemical properties, which can migrate and transform in the environment and exist for a long time [2,3,4]. In addition [2,4], some pollutants, such as polychlorinated biphenyls, polycyclic aromatic hydrocarbons, organophosphorus and organochlorine pesticides, are carcinogenic, teratogenic and mutagenic, accumulating in the human body through the food chain and posing a huge threat to human health [5,6,7]. Therefore, water pollution treatment is becoming more complex and challenging. At present, the commonly used organic pollutant treatment methods include the biological treatment method [8], adsorption method [9,10] and oxidation method [11,12], etc. However, these technologies have diverse limitations of pollutants transfer, incomplete degradation and high energy requirements [13]. For example, the efficiency of biological treatment methods is limited by pollution and easy-to-breed resistant strains [14,15]. Adsorption techniques also have problems recovering and disposing of irreversible adsorbents [13].
Chemical oxidation technology includes direct oxidation and catalytic oxidation. At present, the limitations of direct oxidation cannot be avoided. For example, although ferrate (VI) has a strong oxidizing ability [16], it has poor stability and high pH requirements [17,18]. Hydrogen peroxide (H2O2) is a mild oxidant with limited oxidizing ability, and there are residual and subsequent disposal problems [19,20,21]. Chlorine dioxide needs to be provisionally prepared when used and has high toxicity [12,22,23,24]. Ozone has a remarkable treatment effect but produces bromate, a potential carcinogen, which threatens human health [25,26]. Catalytic oxidation generates active species with strong oxidation, such as hydroxyl radicals (·OH) and sulfate radicals (SO4•−) under external conditions (heating, light, catalysts, etc.) to oxidize organic pollutants. For example [12,19,20,21,22,25,27,28], advanced oxidation processes (AOPs) based on Fenton, ozone and persulfate can remove pollutants more effectively by generating ·OH and SO4•− [27], but they are limited by acidic conditions and high selectivity. In addition, Kosar found that photo-Fenton combined with ozonation showed faster and better efficiency than that of separate oxidation, but the reaction was under pH = 3 [28]. In recent years, non-thermal plasma has also attracted much attention, but high cost and immature technology limit its application and development [29]. Researchers have also studied peracetic acid (PAA) as an emerging oxidant in recent years, which can generate a special alkoxy radical (R-O·) [30,31] with excellent resistance to reaction conditions and has been widely used in foreign countries [32,33].
PAA has been used in sterilization and disinfection and is extensively studied in the field of organic wastewater degradation [33,34,35,36,37]. Compared with traditional chlorine-containing disinfectants, PAA has a stronger oxidizing ability, higher disinfection efficiency and lower toxicity [38,39,40,41,42]. Therefore, PAA is considered an alternative to chlorinated disinfectants. PAA can not only directly oxidize organic compounds such as 17 β-estradiol, 17 α-vinyl estradiol and ibuprofen using its high redox potential [41] but can also be activated by ultraviolet (UV), transition metal ions (such as iron and cobalt ions), transition metal oxides, carbonaceous materials and some composite materials to generate HO·, methyl (CH3·), acetoxy (CH3COO·), acetyl peroxy (CH3C(O)OO·) and some metal complexes to degrade organic pollutants in water and wastewater via its low peroxy bond energy (38 kcal·mol−1) [42,43,44,45]. PAA-based AOPs have been applied to remove phenols, dyes and antibiotics with excellent effect [34,46,47,48,49], but problems such as the unclear activation mechanism, complex preparation process and secondary pollution cannot be ignored. At present, there are few reviews on the degradation of organic pollutants by PAA. AO et al. [32] introduced the mechanism and application of PAA technology for water purification and disinfection in detail but ignored the non-radical degradation pathways. Based on the research of others, this paper introduces the preparation method and characteristics of PAA, summarizes the mechanism and reactivity of the free radical pathway and non-radical pathway for the first time and discusses the main influencing factors. Finally, the shortcomings of the current PAA-based treatment systems are analyzed, and its development is prospected. This review aims to provide a reference for the development of PAA-based technologies that can be practically applied to the treatment of organic pollutants in water.

2. Preparation and Characteristics of PAA

2.1. Characteristics of PAA

PAA is colorless with a pungent acidic smell similar to that of acetic acid. Due to the intramolecular hydrogen bonding between carbonyl oxygen and hydroxyl hydrogen [50,51], as shown in Table 1 and Figure 1a, PAA has a low boiling point (110 °C), a low melting point (0.2 °C), a high pKa (8.2) [52,53] and a five-membered ring [49,52], which makes the neutral form (PAA0) more stable than the anionic form (PAA), existing mainly as a neutral substance even in weakly basic water [53]. Moreover, PAA0 has a higher oxidation capacity than PAA [54]. PAA is unstable and volatile; when the concentration is higher than 15 %, it has the risk of combustion and explosion [32]. In addition, the redox potential of PAA is between 1.06 and 1.96 V, as shown in Table 1, stronger than H2O2 (1.78 V), Cl2 (1.48 V), ClO2 (1.28 V) and FeVI (0.9–1.9 V). Therefore, PAA is widely used in the food processing, healthcare and textile industries as a disinfectant, considered one of the most promising alternatives to chlorine-based disinfectants [55]. Apart from direct disinfection and oxidation, PAA can also produce CH3C(O)O·, CH3C(O)OO· (R-O·) and HO· to improve the disinfection and degradation efficiency by activating [56]. As shown in Figure 1b,c, the stronger polarity of the O-O bond indicated that it is more prone to break. Furthermore, the peroxide bond energy of PAA (159 kJ/mol) is weaker than that of H2O2 (213 kJ/mol), so PAA may exhibit more efficiency [49]. Long-term exposure to high levels of PAA can cause severe irritation to the eyes, skin, mucous membranes and upper respiratory tract [49,54], while low levels may cause occupational asthma [57].
PAA in solution is prone to decomposition and hydrolysis, as shown in Equations (1)–(3), where the hydrolyzed PAA is reduced to acetic acid and H2O2. Da Silva et al. [49] showed that PAA undergoes five stages of hydrolysis in acidic media, and the main factor maintaining the equilibrium between PAA and H2O2 is the small energy difference in each stage. In addition, there are two mechanisms for the spontaneous decomposition of PAA: the acid spontaneously decomposes in response to protonation and PAA attacks on the PAA molecule, forming an active intermediate [49]. Moreover, in the presence of metals, the PAA molecule is catalytically decomposed, as shown in Equation (6). Studies have shown that solutions containing 40% PAA lose 1–2% of the active ingredient per month, with lower concentrations losing more [58,59]. Therefore, we believe that it is essential to determine PAA concentration regularly.
CH 3 C O OOH + H 2 O CH 3 COOH + H 2 O 2 ·
2 CH 3 C O OOH CH 3 COOH + O 2 ·
CH 3 C O OOH CH 3 COOH + O 2 + other   product .

2.2. Synthesis and Detection of PAA

The synthesis of PAA was first investigated by Food Machinery Chemical Co. in the early 20th century [60]. There are two general methods to synthesize PAA: H2O2 oxidation and acetaldehyde oxidation. However, the latter is not commonly applied in practice. H2O2 oxidation involves the reaction of acetic acid (Equation (4)) or acetic anhydride (Equation (5)) with H2O2. However, the reaction of acetic anhydride with H2O2 is difficult to control and may produce diacetyl peroxide as a by-product. In addition, the reaction is exothermic and may cause an explosion. Therefore, acetic acid is the most commonly utilized substance in the preparation of PAA. The reaction of acetic acid with H2O2 is reversible, so researchers add a strongly acidic catalyst to promote the reaction [61]. For example, Saha et al. used a gas diffusion electrode to generate H2O2 in situ for the synthesis of PAA [62].
CH 3 COOH + H 2 O 2 catalyst CH 3 C O OOH + H 2 O .
CH 3 CO 2 O + 2 H 2 O 2 = 2 CH 3 C O OOH + H 2 O .
A commercial PAA solution is a mixture of acetic acid, H2O2, PAA and water [57,61]. Therefore, the amount of H2O2 in PAA solutions should be considered during the study [54,56].
A direct method for the determination of PAA is not yet available. We need to determine the sum of PAA and H2O2 concentrations by iodine titration and the H2O2 concentration in the PAA solution by potassium permanganate titration. The final concentration difference is the concentration of PAA [63,64]. For the remaining PAA concentration in the reaction solution, the N,N-diethyl-p-phenylenediamine colorimetric method is generally used [65]. In addition, Tashkhourian et al. [66] introduced a novel sensor for the simultaneous determination of PAA and H2O2 using silver nanoparticles as a chromogenic reagent.

3. Methods and Mechanisms of Activating PAA

There are two categories according to the main reactants in the PAA activation system: free radical and non-free radical pathways. There are different activation methods for PAA. The free radical pathway can be divided into energy radiation, transition metal catalysis and activated carbon catalysis according to the activation methods. They can generate radicals such as HO·, R-O· and so on, accelerating the degradation of organic pollutants. The non-radical pathway mainly occurs in the transition metal catalysis process, where PAA forms complexes with transition metals. However, there are different mechanisms of different methods.

3.1. Free Radical Pathway

3.1.1. Energy Radiation

As shown in Equation (6), driven by the applied energy radiation, the O-O bond of PAA is cleaved to generate HO· and CH3C(O)O·. This process is the key step of energy radiation, which directly determines the degradation rate of pollutants. In addition, the PAA solution also contains a small amount of H2O2, which can also be activated by energy radiation to generate HO· to participate in the subsequent reaction (Equation (7)).
CH 3 C O OOH energy CH 3 C O O · + HO · .
H 2 O 2 energy 2 HO · .
Subsequently, HO· extracts hydrogen atoms to attack PAA to generate radicals, of which CH3C(O)OO· is the main one, as shown in Equations (8)–(10).
CH 3 C O OOH + HO · CH 3 C O OO · + H 2 O .
CH 3 C O OOH + HO · CH 3 CO · + O 2 + H 2 O .
CH 3 C O OOH + HO · CH 3 COOH + · OOH .
CH3C(O)O can be turned into a methyl radical (CH3·) and carbon dioxide (CO2) by a decarboxylation reaction (Equation (11)). Additionally, CH3· can rapidly react with oxygen to form peroxy radical (CH3OO·) (Equation (12)), followed by CH3OO· attacking PAA to form CH3C(O)OO· (Equation (13)). Meanwhile, CH3OO· can also undergo bimolecular decay by itself (Equations (14)–(16)), while CH3C(O)OO· and CH3C(O)O· may undergo radical–radical coupling (Equations (17) and (18)).
CH 3 C O O · CH 3 · + CO 2 ·
CH 3 · + O 2 CH 3 OO · .
CH 3 C O OOH +   CH 3 C O O · CH 3 C O OO · +   CH 3 COOH .
CH 3 OO · + CH 3 OO · HCHO + CH 3 OH + O 2 ·
CH 3 OO · + CH 3 OO · 2 HCHO + H 2 O 2 ·
CH 3 OO · + CH 3 OO · 2   CH 3 CO · + O 2 ·
CH 3 C O OO · + CH 3 C O OO · 2   CH 3 C O O · + O 2 ·
CH 3 C O O · + CH 3 C O O · CH 3 C O O 2 ·
The above reaction also involves HO· reacting with CH3C(O)O· to form PAA (Equation (19)), forming a cyclic reaction system.
CH 3 C O O · + HO · CH 3 COOOH .
H2O2 in PAA solution can produce other radicals besides HO·, making the energy activation system of PAA more complex. Zhang et al. proposed a reaction mechanism of UV/PAA in the presence of H2O2 and acetate ions [67] (Figure 2), where a high steady-state concentration of superoxide radicals, superoxide radical anions and ·OOCH2C(O)O was observed.

3.1.2. Transition Metal Catalysis

There are two types of transition metal catalysis: homogeneous and heterogeneous catalysis, both of which have similar activation principles. As shown in Figure 3, unlike the energy activation mechanism, transition metals can both gain and lose electrons, resulting in a cyclic activation mechanism (Equations (20) and (21)).
X n + +   CH 3 C O OOH X n + 1 + + CH 3 C O O · + OH .
X n + 1 + +   CH 3 C O OOH X n + +   CH 3 C O OO · + H + .
The current transition metal catalysts for PAA are mainly based on Co, Fe and Mn. Taking Fe as an example, we introduced the PAA activation mechanism in detail. In Equations (22)–(25), R-O· radicals and CH3C(O)O, FeIVO2+, HO· and CH3COOH are generated. Subsequently, FeIVO2+ continues to react with PAA or H2O to form Fe3+, which participates in the reaction of Equations (26) and (27). Reactions of Equations (8)–(19) are still present in the process, so lots of free radicals and other products are formed.
Fe 2 + +   CH 3 C O OOH Fe 3 + +   CH 3 C O O · +   OH .
Fe 3 + +   CH 3 C O OOH CH 3 C O OO · + Fe 2 + + H + .
Fe 2 + +   CH 3 C O OOH Fe 3 + +   CH 3 C O O +   HO · .
Fe 2 + +   CH 3 C O OOH Fe IV O 2 + +   CH 3 COOH .
2 Fe IV O 2 + +   CH 3 C O OOH CH 3 C O O + 2 Fe 3 + +   OH + O 2 ·
Fe IV O 2 + + H 2 O HO · + Fe 3 + +   OH .
Under acidic conditions, H2O2 in PAA solution also experiences the reaction of Equations (28)–(31) [54], which is known as the Fenton reaction. The Fenton reaction is widely used for wastewater treatment.
Fe 2 + + H 2 O 2 HO · + Fe 3 + +   OH .
Fe 2 + + H 2 O 2 Fe IV O 2 + + H 2 O .
Fe 3 + + H 2 O 2 HO 2 · + Fe 2 + + H + .
Fe IV O 2 + + H 2 O 2 HO 2 · + Fe 3 + + OH .

3.1.3. Activated Carbon Catalysis

Activated carbon fiber (ACF) has a high specific surface area, a large number of active sites, functional groups and nonbonding free electrons. Unpaired free electrons in ACF may be transferred to PAA, promoting the formation of free radicals. Similar to energy radiation, as shown in Figure 4, Zhou et al. used density functional theory to simulate the reaction process and found that the addition of ACF grew the length of the O-O bond of PAA, resulting in the O-O bond breaking more easily and generating CH3C(O)OO· and HO· radicals [48].

3.2. Non-Free Radical Pathway

In the system of activating PAA by transition metal, it is widely believed that R-O· is the main reactive substance. There are many researchers who have largely used radical quenching experiments combined with electron paramagnetic resonance radical detection pathways to determine the major reactive substance in transition metal/PAA systems, which is controversial as it depends on the dominant reaction of the free radical. However, research has shown that when transition metal activates persulfate, high-valent metals play a major role rather than free radicals [31]. Zhao et al. [68] also found that the oxygen atom transfer pathway plays a primary role in the activation of PAA by Co. As shown in Figure 5a, the oxygen atom coordinates with Co(II) to form Co(II)-OO(O)CCH3, which plays a key role in the degradation of organic pollutants, and Co(III) and R-O· are the secondary reaction species. In addition, as shown in Figure 5b, Liu et al. [44] pointed out that Co(IV) acts as an oxidant rather than active free radicals in the Co(II)-doped G-C3N4-PAA system. They proposed two catalytic hypotheses: the atom Co is coordinated to an atom O near the atom H, and the atom O is coordinated to a peroxide bond near the atom C. The positively charged Co in the first hypothesis acts as an electron donor, decreases the activity of PAA and lengthens the O-O band, triggering the spontaneous dissociation of PAA. The second hypothesis of PAA acting as an electron donor allows electrons to be transferred from PAA to the electron-deficient region, forming surface-bound CH3C(O)OO·. Furthermore, Manoli et al. used the Fe(VI)-PAA process to degrade carbamazepine and found that the most dominant active species in the system was Fe(VI) with CH3C(O)OO reacting to produce Fe(V)/Fe(IV) as shown in Equations (32) and (33) [69].
Fe VI O 4 2 + CH 3 C O OO Fe V O 4 3 + CH 3 C O OO · .
Fe VI O 4 2 + CH 3 C O OO Fe IV O 4 2 + CH 3 C O O + O 2 .

4. Reactivity of PAA-Based AOPs

4.1. Oxidation of PAA Alone

PAA has been used as a broad-spectrum, highly effective disinfectant in the food industry, medical devices/pharmaceuticals, textiles and water treatment processes. PAA is selective when used alone. As shown in Table 2, the degradation rates of different pollutants with PAA under the same conditions were found to be different. For example, mefenamic acid (≈90%) and ibuprofen (≈30%), owing to the acetyloxy radical, showed high reactivity to electron-rich functional groups. Several studies have shown that water quality also plays a major role in the degradation of pollutants by PAA [70,71,72]. Wang et al. found that there were differences in the degradation effects of the same pollutant under different water quality conditions by PAA, which were attributed to PAA generally reacting more uniformly with the pharmaceutical ingredients from wastewater containing low COD [72]. In addition, the dosage of PAA also influences the oxidation effect of PAA. Concentrations of 50 μg·L−1 17 β-estradiol and 17 α-ethinyl estradiol were able to be completely removed by 15 mg·L−1 PAA, whereas 50 μg·L−1 Brilliant Red X-3B was removed by less than 5% with 380 mg·L−1 mM PAA [41,48]. Furthermore, as shown in Table 2, many studies have shown that the removal efficiencies of PAA for other organic pollutants are also less than 20%. Therefore, PAA alone was not effective in degrading pollutants.

4.2. PAA-Based AOPs

4.2.1. Homogeneous Catalysis

Catalysis in the homogeneous phase involves a homogeneous mixture of catalysts and reactants. The common homogeneous catalyst of PAA includes metal ions and inorganic anions. In Table 3, the transition metal ions that can be used to activate PAA are mainly Fe2+, Cu2+, Co2+, Ru3+ and Mn2+. Among them, Co2+ has received wide attention due to its remarkable effect. Co2+ showed superior removal in less time with fewer reagents and lower pH selectivity than Fe2+ and Cu2+ in removing bisphenol-A [72,77]. Kim et al. [77] also investigated the degradation of naproxen, sulfamethoxazole (SMX) and carbamazepine by the Co2+/PAA process and confirmed the excellent performance of the process. However, some researchers found that the Co2+/PAA process is still selective for pollutants. Zhao et al. [68] investigated the removal of nitrobenzene by the Co2+/PAA process under the same conditions and found that the degradation rate was only 18.8%. In addition, there are still some controversies regarding the interpretation of the oxidation mechanism of the Co2+/PAA process. Wang et al. [56] showed that R-O· is the main species in the process of Co(II)/PAA, while Zhao et al. [68] found the complex Co(II)-OO(O)CCH3 with a stronger oxidizing ability to remove pollutants.
In contrast to toxic Co(II) and Mn(II), Fe(II) and Cu(II) are often considered ideal transition metal catalysts. However, the reaction is largely stopped when Fe(II) translate to Fe(III) [78]. Cu(II) is more soluble and less prone to generate sludge, but it is ineffective because of its lower reduction potential (ECu(II)/Cu(I) = 0.167 V) [79]. In order to improve the Cu (II)/PAA, Wang [80] et al. added HCO3−(CO32−) to the Cu(II)/PAA process to form a highly reactive complex Cu(II)-HCO3(CO32−). Instead of Cu(II), zero-valent copper has also been studied to activate PAA by generating low-valent copper ions (Cu(I)) under acidic conditions [81].
Furthermore, some scholars also used inorganic anions to activate PAA. Deng et al. [42] demonstrated that phosphate could activate PAA to produce HO· and R-O· by electron paramagnetic resonance and free radical scavenging experiments. Wang et al. [82] found that the Cl /PAA process removes Rhodamine B by producing R-O· and 1O2 and showed high removal efficiency for Rhodamine B over a wide pH range.
Table 3. Studies of activating PAA in homogeneous phase system.
Table 3. Studies of activating PAA in homogeneous phase system.
CatalystsCCatalyst (μM)ContaminantsC0 (μM)CPAA (μM)pHTime
(min)
Max. Removal
Efficiency (%)
References
Fe(VI)200SMX101009.0 ± 0.11~100[69]
Sulfadimethoxine~100
Trimethoprim>80
Atenolol>85
Propranolol>90
Caffeine85
Fe(II)100Methylene blue151003.0~8.2120~90[54]
Naproxen~100
Bisphenol-A>80
Fe(II)0.4 mMBisphenol-A60 mg·L−140 mg·L−13.510100[72]
Fe(II)0.4 g·L−1polyacrylamide200 mg·L−110 mg·L−1315>80[83]
Fe(II)5 mg·L−1Diclofenac11007180[84]
Cu(II)0.4 mMBisphenol-A60 mg·L−120 mg·L−13.560>95[72]
Cu(II)-HCO3(CO32−)15Diclofenac1609.32085[80]
Co(II)15Bisphenol-A151003.0~8.130100[77]
Naproxen100
SMX98.5
Carbamazepine87.7
Co(II)0.8 μMSMX10100715~90[56]
Co(II)0.01 mMAcid orange 70.05 mM0.2 mM76092[85]
Methylene blue52
Congo red98
Crystal Violet67
Co(II)2.0Bisphenol-A8.010072093.6[68]
SMX99.3
Phenol66.2
Nitrobenzene18.8
Mn(II)0.1~1 mg·L−1Orange II17.5 mg·L−138~3802 mg·L−19.5n.r.n.r.[86]
phosphate0.1 MDiclofenac5 μM0.55 mM7.445~100[42]
Cl400 mMRhodamine B10 mg·L−12.0 mMn.r.10~100[82]

4.2.2. Heterogeneous Catalysis

  • Energy Catalysis
PAA can be activated by electric current, heat, microwave, ultraviolet light or sunlight. However, UV light is the most popular. Table 4 collect the results of different energy activations of PAA for the degradation of contaminants. Hollman et al. [36] used UV/PAA to produce R-O· to remove SMX, fluoxetine, carbamazepine and naproxen completely within 30 min under neutral conditions. However, electrochemistry and natural light need more time to remove pollutants effectively. The degradation rate of methylene blue under electrochemistry was reported to be 93.99% in 120 min [87], and venlafaxine was removed at 98.5% under natural light with 40 °C heating in 60 min [88]. PAA also releases more active substances when heated. Wang et al. [89] found that SMX was completely removed in 25 min under 60 °C/PAA. Energy activation has significant effects, but its high operating costs and low safety factor limit its further development.
The factors influencing the degradation effect of the energy/PAA process, such as energy intensity, pH, water quality, and PAA concentration on pollutants, also deserve to be studied [90]. The removal rate of the target pollutants generally increases and then decrease as the PAA concentration increases. Studies have shown that higher PAA concentrations improve the removal of organic pollutants by UV/PAA technology [63,64,91]. However, Hollman et al. and Li et al. [36,92] also found that too many dosages of PAA show a certain quenching effect on HO·. For the effect of pH, some studies have shown that PAA has a stronger HO· scavenging effect in an alkaline environment, and the photolysis rate is higher than that of PAA0 [63]. Therefore, the higher the pH, the worse degradation. It can be concluded in Table 3 that the degradation rate is higher under neutral or acidic conditions. However, Daswat et al. found that alkaline conditions (optimum pH = 11) are most suitable for the UV/PAA process to treat chlorophenol industrial wastewater [90], which is probably attributed to the complex water quality conditions of industrial wastewater. Therefore, some coexisting substances in the water also have an impact on the degradation of organic pollutants. Chen et al. investigated the effect of CO32/HCO3, HA and Cl common in water on the degradation of naphthalene (NAP) by the UV/PAA process and found that the effect of CO32/HCO3 and Cl was negligible and HA showed a significant inhibitory effect [64]. In addition, NO3, as a photosensitive substance, has been shown to produce HO· under UV light, but the product NO2 is a strong HO· bursting agent and shows a slight increase in the overall degradation rate [93].
Table 4. Studies of activating PAA by energy.
Table 4. Studies of activating PAA by energy.
EnergyContaminantsC0CPAApHTime (min)Temperature (°C)Max. Removal Efficiency (%)Free RadicalsReferences
electrochemical oxidationmethylene blue10 mg·L−13.6 mmol·L−1312025 ± 293.99HO•>R−O• [87]
HeatSMX5 μM0.025–0.2 mM4–92560~100R−O• [89]
microwavethiophene sulfur0.10%0.50%n.r.3n.r.12.07n.r. [94]
solar radiationVenlafaxine5 mg·L−128.6 mmol −12.3604098.5n.r. [88]
UV254SMX5 mg·L−15~100 mg·L−17 ± 0.230n.r.~100R−O• [36]
Fluoxetine~100R−O•
Carbamazepine~100R−O•
Naproxen~100R−O•
UV254methylene blue16 µM20 mg·L−171420 ± 1~80n.r. [64]
UV254Bezafibrate1 µM1 mg·L−17.1120n.r.>90HO•>R−O• [63]
Carbamazepine30>90
Clofibric acid10>90
Diclofenac<5>90
Ibuprofen30>90
Naproxen10>95
UV254methylene blue10 mg·L−150 mg·L−17.60–120n.r.85%n.r. [91]
2.
Catalyst Catalysis
Although the activation of PAA by metal ions has the advantages of low energy consumption and high efficiency, the biological toxicity they bring will cause secondary pollution to the environment, which limits their practical application. To overcome this drawback, researchers have investigated heterogeneous catalysts. In Table 5, the main heterogeneous catalysts for PAA are metal oxides, activated carbons and metal–carbon materials. Non-metallic carbonaceous catalysts have been extensively studied due to their large specific surfaces, lack of secondary contaminants and excellent chemical properties [95,96,97]. Water treatment processes have extensively investigated the use of carbon-based materials for activating peroxides. However, only one study has reported the use of carbon catalysts to activate PAA for the degradation of organic pollutants. Zhou et al. [48] investigated that 92.5% reactive brilliant red X-3B was degraded in 45 min under neutral conditions by ACF/PAA. In contrast, metal-based heterogeneous catalysts are more extensively studied and perform better. It has been reported that SMX and 2,4-dichloropheno can be removed completely by adding LaCoO3 or Co@MXenes and PAA to the solution [55,58,74,77]. Co3O4 exhibited similar properties to LaCoO3 properties [75]. However, metal-based catalysts loaded on carbon materials are more easily recovered. Wang et al. [98] activated PAA by Fe(II)-zeolite instead of Fe(II), which exhibited a solid–liquid interface reaction. The process can effectively degrade organic pollutants under neutral conditions and control the generation of iron sludge. Therefore, the heterogeneous catalysts not only have an excellent effect on the activation of PAA but also have the advantages of low toxicity and recyclability, which is an excellent catalyst method for the application of PAA in the treatment of wastewater.

5. The Factors That Influence the Removal of Organic Pollutants in PAA-Based Processes

Through the above discussion, we learned that there are many ways to activate PAA and the PAA-based AOPs have strong oxidative properties. However, some factors, including PAA dosage, catalysts dosage, pH, temperature and water quality components, affect the degradation effect of the PAA system.

5.1. PAA Dosage

Generally, more oxidants can improve degradation efficiency. As the PAA concentration increases, more reactive radicals will be generated in the free radical pathway, but too much PAA will prevent it. Zhang et al. [55] found that with the increase in PAA concentration, the free radicals in the Co@MXenes/PAA process increased so that the reaction constant and degradation rate of 2,4-dichloropheno increased. Nevertheless, 2,4-dichloropheno degradation was inhibited when PAA concentrations exceeded 1.04 mM. Similarly, Kim et al. [54] also found that the degradation of methylene blue by the Fe(II)/PAA process was inhibited when the PAA concentration was increased to 1 mM. This may be attributed to the free radical scavenging effect of excess PAA. In addition, there are H2O2 in the PAA solution, which shows a scavenging effect on R-O· (Equations (34)–(36)). Furthermore, excess PAA absorbs UV light, reducing the number of target compounds that are directly photolyzed [104]. However, the effect of PAA dosage on non-radical reactions has not been studied by scholars.
H 2 O 2 + CH 3 C O OO · HO 2 · + CH 3 C O OOH .
H 2 O 2 + CH 3 COO · HO 2 · + CH 3 COOH .
O 2 · +   CH 3 C O OO · O 2 + CH 3 C O OO .

5.2. Catalysts Dosage

Under the free radical pathway, the catalysts will promote the generation of more active radicals with the increase of the dosage. Kim et al. [77] found that increasing Co in the Co/PAA process accelerates the decomposition of PAA. Wang et al. [56] showed that as the initial concentration of Co increased from 0.2 to 3.2 μM, the number of active radicals generated increased, and the degradative efficiency of SMX increased from 61.2% to nearly 100% within 15 min, which proved that Co3+/Co2+ transformation occurred in Co/PAA through the redox cycle of Equations (37) and (38). Increased catalyst provides more active sites in heterogeneous catalysis. According to Zhang et al. [55], when Co@MXene was added to the reaction, the catalytic sites were increased, which speeds up the breakdown of 2,4-dichloropheno. The photocatalytic process also has a similar rule. By conducting experiments at different UV intensities of 0.65, 1.87 and 3.50 kW/m3, Hollman et al. [36] found that the increase in UV intensity accelerates the degradation of pollution.
Co 2 + + CH 3 CO 3 H Co 3 + + CH 3 CO 2 · + OH .
Co 3 + + CH 3 CO 3 H Co 2 + + CH 3 CO 3 · + H + .
However, due to the limitation of oxidants in the system, too much catalyst cannot increase the degradation efficiency indefinitely and may also have a scavenging effect on free radicals. Kim et al. [54] found that Fe(II) dosage did not significantly affect the degradation rate of methylene blue in the initial stage but affected the second reaction stage.

5.3. pH

The pH of the reaction is one of the most important factors affecting the performance of PAA-based AOPs. First, pH affects the acid–base balance of PAA, which in turn affects the generation of free radicals. The pKa value of PAA was 8.2, and the PAA0 was dominant at pH 3–7 (neutral species fraction fPAA0 = 1~0.94). As the pH approached 8.2, fPAA0 decreased to 0.56, which suggested that under alkaline conditions, reactive radicals such as ·OH may be suppressed by their reduced redox potentials and competition with other compounds [104]. PAA is a weaker oxidant than PAA0 but reacts more readily with some radical species (·OH reacts with PAA and PAA0 with second-order reaction rate constants of 9.97 × 109 and 9.33 × 108 M−1·s−1, respectively). Kim et al. [54] investigated the catalytic effect of Fe2+ on PAA and found that at initial pH ≤ 6.0, the PAA content decreased rapidly within 2 s while the pH value increased to 8.1, and the catalytic effect of Fe2+ on PAA was poor. In contrast, the molar absorption coefficient of PAA and PAA0 at 254 nm was higher for the photocatalytic process, resulting in a faster photolysis rate of PAA and PAA0. For example, Mukhopadhyay and Daswat [105] reported that the optimum pH for the degradation of 4-chlorophenol by the UV/PAA process was 9.5. The pH value also has an effect on the morphology of the metal. Fe(II) forms hydroxyl complexes at higher pH, and the catalytic activity is inhibited.
Fe2+ is oxidized by oxygen more readily as pH increases, decreasing the availability of Fe2+ reacting with PAA (or H2O2) at higher pH, while the Fe(III) generated during the reaction was precipitated at lower pH due to the complexation of hydroxide. In addition, pH also affects the surface charge of metal oxides during the heterogeneous activation of PAA. Zhang et al. [55] investigated the degradation of 2,4-dichloropheno by the CO@MXenes/PAA process and found that under neutral conditions, 2,4-dichloropheno degraded faster. Furthermore, the activity of PAA and the surface properties of the catalyst are affected by pH. Under alkaline conditions, Co@MXenes has a negative charge, and there are mainly conjugated bases of PAA in solution, which are attracted or suppressed by electrostatic repulsion. PAA is hydrolyzed more rapidly, and inactive cobalt hydroxide complexes are formed on the surface of Co@MXenes, reducing the efficiency of the system. Under acidic conditions, PAA mainly exists in the form of PAA0, which is much more stable than PAA, inhibiting the generation of active free radicals. Similar conclusions remain in the non-radical pathway. Zhao et al. [68] found in the experiment of bisphenol A degraded by Co(II)/PAA that as the pH increased from 3.0 to 7.0, the degradation efficiency of bisphenol A gradually decreased, which is probably attributed to the protonation state of the Co(II)-OO(O)CCH3 species was changed, while the fraction of Co2+ decreased from 98.7% to 74.1% with the pH increased from 7.0 to 9.0, and there was no Co2+ species but Co(OH)2 with less reactive produced when pH reached 11.0.

5.4. Temperature

In practical applications, temperature is also an important factor affecting the efficiency of PAA-based AOPs. Zhang et al. [55] investigated the temperature dependence of 2,4-dichloropheno degradation in the Co@MXenes/PAA process and found that 2,4-dichloropheno was removed by 93.9% within 20 min at 10 °C, but the degradation rate reached 100% within 15 min when the temperature increased to 30 °C, and kobs increased from 0.1714 min−1 increased to 0.3485 min−1. The activation energy of 2,4-dichloropheno was 43.4 kJ/mol, lower than the reported Co(OH)F@MXenes/PMS system (57.32 kJ/mol) and Co3O4@MXenes/PMS system (54.4 kJ/mol), which indicates that the reaction energy barrier is lower in the Co@MXenes/PAA system.

5.5. Water Quality Components

There are various natural organic matters and dissolved inorganic anions in water and wastewater. Due to the special structures and properties of these substances, they become factors that cannot be ignored in the PAA process.

5.5.1. Anions

HCO3/CO32 and Cl are common inorganic anions in water. They can affect the treatment efficiency of PAA activation systems by competing with organic pollutants for free radicals. As shown in Equations (39)–(43) [46,82], the reaction of Cl and HO· will generate HClO·, which is mainly converted to HO· rather than Cl· under neutral conditions. At the same time, Cl has the ability to react with R-O· to create Cl·, which is converted to HO·. When the concentration of Cl is very high, chlorine radicals including Cl·, Cl2· and OCl· are generated in the system. In contrast, when the concentration of Cl is low, Cl· reacts with H2O/OH to generate HOCl·, which is easily converted to HO· under acidic conditions. In addition, Cl· and Cl2· will further generate Cl2 and HOCl. In spite of the high redox potentials (E0(Cl·/Cl) =2.5V, E0(Cl2·/Cl) = 2.2 V), these active chlorine species have stronger specificity. Therefore, when the target pollutants can be readily oxidized by active chlorine species, Cl will improve the treatment efficiency in the PAA-based AOPs. Moreover, HOCl tends to be the dominant active species due to its longer lifetime than R-O· and HO· under high concentrations of Cl. Wang et al. [56] investigated that the addition of Cl has minimal effects on the degradation of SMX in the Co/PAA system, showing that Cl can scavenge R-O· as a by-product of SMX decomposition (Equation (16)). Wu et al. [75] observed that Cl significantly inhibited the degradation of orange G in the Co3O4/PAA system and attributed it to the low reactivity of the generated active chlorine to orange G. Strong oxidizing properties of chlorine species coming from PAA-based AOPs have high selectivity, so it is confirmed that the reactions produced by Cl affect the degradation of different organic pollutants in different ways.
CH 3 CO 3 · + Cl + H + Cl · + CH 3 CO 3 H .
Cl + HO · HClO · .
HClO · + H + Cl · + H 2 O .
Cl +   Cl · Cl 2 .
Cl · + Cl · Cl 2 .
The alkalinity of water is mostly affected by HCO3/CO32−, which are known to scavenge free radicals. As shown in Equations (44) and (45), CO32− and HCO3 can quench ·OH, resulting in less reactive CO3· [46,106]. At the same time, the combination of HCO3 and metal ions may form complexes. For example, Wang et al. [56] examined the effect of HCO3/CO32− on the degradation of SMX by the Co/PAA process and found that though SMX degradation efficiency decreased with the increase of HCO3/CO32− concentration, the difference between HCO3 and R-O· persisted. The Co/PAA reaction has a very low-rate constant, so the inability of this reaction would not primarily be due to the consumption of R-O· through HCO3. However, HCO3 will form non-reactive Co2+–HCO3 complexes during the reaction, which limits the activation of PAA and inhibits the degradation of SMX. Wang et al. [47] studied the degradation process of SMX in a CoFe2O4/PAA system and found that the residual PAA content in the system with 2 mM HCO3 was higher than that without addition.
HCO 3 + HO · CO 3 · + H 2 O .
CO 3 2 + HO · CO 3 · + OH .

5.5.2. Dissolved Organic Matter (DOM)

DOM reacts with R-O· and HO· (mainly R-O·) to inhibit the degradation efficiency of PAA-based AOPs, and the higher the DOM concentration, the more obvious the inhibition. Wang et al. [56] showed that when the HA concentration of the CoFe2O4/PAA system was increased from 0 to 10 mg·L−1, the SMX removal rate decreased from 74.7% to 5.8%, and the kobs decreased from 0.047 to 0.002 min−1 in 30 min. In addition to adsorbing on the catalyst, some HA groups, such as phenolic hydroxyl and carboxyl groups, would inhibit PAA activation [55]. Furthermore, HA may also scavenge free radicals. Zhao et al. [68] evaluated the removal of bisphenol A in the Co(II)/PAA process and found that natural organic matter (NOM) significantly inhibited the removal efficiency, but PAA decomposition was only slightly inhibited, indicating that inhibition may be due to the competition between NOM and bisphenol A.

6. Conclusions and Future Prospects

The latest research on PAA-based AOPs for the removal of organic pollutants was reviewed in this paper. The methods principle and reactivity of activating PAA were discussed. Currently, the most prominent activation methods are energy radiation (UV, solar radiation, ultrasound, heating), transition metal and activated carbon material catalysis. Most studies showed that PAA attacked pollutants mainly by generating strong oxidative free radicals (R-O·, HO·), and R-O· is dominant in most of the reactions. In recent years, some scholars have pointed out a non-radical pathway, which removes pollutants by generating complexes or high-valent metal species, and confirmed its key role in the degradation process. The removal effect of PAA-based AOPs is affected by various factors, such as PAA/catalyst dosage, pH, temperature and water quality components (inorganic anions and DOM). PAA-based AOPs have gradually become a hot spot in the research of removing organic pollutants due to their excellent degradation effect.
However, PAA-based AOPs still face some challenges in removing organic pollutants. First, the activation mechanism of PAA is still controversial, especially since the emergence of non-radical pathways in recent years has broken people’s potential awareness of radical pathways. Secondly, the reaction mechanisms, rate constants and identification methods of organic radicals (R-O·) and non-radicals have not been systematically studied. Furthermore, transition metal catalysis will lead to metal leakage. Additionally, the toxicity assessment of by-products released from PAA activation of organic pollutants is an important research project, but there are few reports on it. Finally, the existing research is mainly aimed at small-scale laboratory experiments, and the technical development of its practical application is incomplete.

Author Contributions

Writing—original draft preparation, C.S. and Y.W.; writing—review and editing, C.S. and J.G. and S.B. and Y.Z. and N.M.; supervision, C.L.; project administration, C.L.; funding acquisition, C.L. All authors have read and agreed to the published version of the manuscript.

Funding

This work was founded by the Shanghai Natural Science Foundation, grant number No. 20ZR1438200 and the National Natural Science Foundation of China, grant number No. 51778565.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Ling, M.; Lv, C.; Guo, X. Quantification method of water environmental value loss caused by water pollution based on emergy theory. Desalination Water Treat. 2018, 129, 299–303. [Google Scholar] [CrossRef] [Green Version]
  2. Liu, D.; Yuan, Y.; Wei, Y.; Zhang, H.; Si, Y.; Zhang, F. Removal of refractory organics and heavy metals in landfill leachate concentrate by peroxi-coagulation process. J. Environ. Sci. 2022, 116, 43–51. [Google Scholar] [CrossRef] [PubMed]
  3. Ply, A.; Wjz, A.; Lhk, A.; Rfs, B.; Xjg, A.; Xi, Y.A.; Yan, C.A.; Wzl, A. Endow activated carbon with the ability to activate peroxymono-sulfate for the treatment of refractory organics. Green Chem. Eng. 2022. [Google Scholar] [CrossRef]
  4. Tawfik, A.; Alalm, M.G.; Awad, H.M.; Islam, M.; Qyyum, M.A.; Al-Muhtaseb, A.H.; Osman, A.I.; Lee, M. Solar photo-oxidation of recalcitrant industrial wastewater: A review. Environ. Chem. Lett. 2022, 20, 1839–1862. [Google Scholar] [CrossRef]
  5. Da Silva, F.C., Jr.; Felipe, M.B.M.C.; de Castro, D.E.F.; Araújo, S.C.D.S.; Sisenando, H.C.N.; de Medeiros, S.R.B. A look beyond the priority: A systematic review of the genotoxic, mutagenic, and carcinogenic endpoints of non-priority PAHs. Environ. Pollut. 2021, 278, 116838. [Google Scholar] [CrossRef]
  6. Sun, S.; Shen, J.; Li, D.; Li, B.; Sun, X.; Ma, L.; Qi, H. A new insight into the ARG association with antibiotics and non-antibiotic agents—Antibiotic resistance and toxicity. Environ. Pollut. 2021, 293, 118524. [Google Scholar] [CrossRef] [PubMed]
  7. Yang, H.; Wang, X. Mechanism of removal of pharmaceuticals and personal care products by nanofiltration membranes. Desalination Water Treat. 2015, 53, 2816–2824. [Google Scholar] [CrossRef]
  8. Xu, R.; Fang, F.; Wang, L.; Luo, J.; Cao, J. Insight into the interaction between trimethoprim and soluble microbial products produced from biological wastewater treatment processes. J. Environ. Sci. 2022, 124, 130–138. [Google Scholar] [CrossRef]
  9. Lobato-Peralta, D.R.; Duque-Brito, E.; Ayala-Cortés, A.; Arias, D.; Longoria, A.; Cuentas-Gallegos, A.K.; Sebastian, P.; Okoye, P.U. Advances in activated carbon modification, surface heteroatom configuration, reactor strategies, and regeneration methods for enhanced wastewater treatment. J. Environ. Chem. Eng. 2021, 9, 105626. [Google Scholar] [CrossRef]
  10. Wang, T.; He, J.; Lu, J.; Zhou, Y.; Wang, Z.; Zhou, Y. Adsorptive removal of PPCPs from aqueous solution using carbon-based composites: A review. Chin. Chem. Lett. 2021, 33, 3585–3593. [Google Scholar] [CrossRef]
  11. Dong, F.; Li, C.; Ma, X.; Lin, Q.; He, G.; Chu, S. Degradation of estriol by chlorination in a pilot-scale water distribution system: Kinetics, pathway and DFT studies. Chem. Eng. J. 2020, 383, 123187. [Google Scholar] [CrossRef]
  12. Gan, W.; Ge, Y.; Zhu, H.; Huang, H.; Yang, X. ClO2 pre-oxidation changes the yields and formation pathways of chloroform and chloral hydrate from phenolic precursors during chlorination. Water Res. 2019, 148, 250–260. [Google Scholar] [CrossRef]
  13. Zheng, H.; Hou, Y.; Li, S.; Ma, J.; Nan, J.; Li, T. Recent advances in the application of metal organic frameworks using in advanced oxidation progresses for pollutants degradation. Chin. Chem. Lett. 2022. [Google Scholar] [CrossRef]
  14. Lin, Q.; Tan, X.; Almatrafi, E.; Yang, Y.; Wang, W.; Luo, H.; Qin, F.; Zhou, C.; Zeng, G.; Zhang, C. Effects of biochar-based materials on the bioavailability of soil organic pollutants and their biological impacts. Sci. Total Environ. 2022, 826, 153956. [Google Scholar] [CrossRef]
  15. Yang, L.; Zhou, Y.; Chen, L.; Chen, H.; Liu, W.; Zheng, W.; Andersen, M.E.; Zhang, Y.; Hu, Y.; Crabbe, M.J.C.; et al. Single enrichment systems possibly underestimate both exposures and biological effects of organic pollutants from drinking water. Chemosphere 2022, 292, 133496. [Google Scholar] [CrossRef]
  16. Zhang, T.; Dong, F.; Luo, F.; Li, C. Degradation of sulfonamides and formation of trihalomethanes by chlo-rination after pre-oxidation with Fe(VI). J. Environ. Sci. 2018, 73, 89–95. [Google Scholar] [CrossRef]
  17. Shi, Z.; Wang, D.; Gao, Z.; Ji, X.; Zhang, J.; Jin, C. Enhanced ferrate oxidation of organic pollutants in the presence of Cu(II) Ion. J. Hazard. Mater. 2022, 433, 128772. [Google Scholar] [CrossRef]
  18. Tian, S.-Q.; Wang, L.; Liu, Y.-L.; Ma, J. Degradation of organic pollutants by ferrate/biochar: Enhanced formation of strong intermediate oxidative iron species. Water Res. 2020, 183, 116054. [Google Scholar] [CrossRef]
  19. Rahmah, A.U.; Harimurti, S.; Murugesan, T. Experimental investigation on the effect of wastewater matrix on oxytetracycline mineralization using UV/H2O2 system. Int. J. Environ. Sci. Technol. 2017, 14, 1225–1233. [Google Scholar] [CrossRef]
  20. Chaves, F.P.; Gomes, G.; Della-Flora, A.; Dallegrave, A.; Sirtori, C.; Saggioro, E.M.; Bila, D.M. Comparative endocrine disrupting compound removal from real wastewater by UV/Cl and UV/H2O2: Effect of pH, estrogenic activity, transformation products and toxicity. Sci. Total Environ. 2020, 746, 141041. [Google Scholar] [CrossRef]
  21. Qu, C.; Liang, D.-W. Novel electrochemical advanced oxidation processes with H2O2 generation cathode for water treatment: A review. J. Environ. Chem. Eng. 2022, 10, 107896. [Google Scholar] [CrossRef]
  22. He, G.; Zhang, T.; Zheng, F.; Li, C.; Zhang, Q.; Dong, F.; Huang, Y. Reaction of fleroxacin with chlorine and chlorine dioxide in drinking water distribution systems: Kinetics, transformation mechanisms and toxicity evaluations. Chem. Eng. J. 2019, 374, 1191–1203. [Google Scholar] [CrossRef]
  23. Li, C.; Luo, F.; Duan, H.; Dong, F.; Chen, X.; Feng, M.; Zhang, Z.; Cizmas, L.; Sharma, V.K. Degradation of chloramphenicol by chlorine and chlorine dioxide in a pilot-scale water distribution system. Sep. Purif. Technol. 2019, 211, 564–570. [Google Scholar] [CrossRef]
  24. Su, R.; Huang, L.; Li, N.; Li, L.; Jin, B.; Zhou, W.; Gao, B.; Yue, Q.; Li, Q. Chlorine dioxide radicals triggered by chlorite under visible-light irradiation for enhanced degradation and detoxification of norfloxacin antibiotic: Radical mechanism and toxicity evaluation. Chem. Eng. J. 2021, 414, 128768. [Google Scholar] [CrossRef]
  25. Pearce, R.; Hogard, S.; Buehlmann, P.; Salazar-Benites, G.; Wilson, C.; Bott, C. Evaluation of preformed monochloramine for bromate control in ozonation for potable reuse. Water Res. 2022, 211, 118049. [Google Scholar] [CrossRef]
  26. Wang, Y.; Man, T.; Zhang, R.; Yan, X.; Wang, S.; Zhang, M.; Wang, P.; Ren, L.; Yu, J.; Li, C. Effects of organic matter, ammonia, bromide, and hydrogen peroxide on bromate formation during water ozonation. Chemosphere 2021, 285, 131352. [Google Scholar] [CrossRef]
  27. Zhang, K.; Zhou, X.; Zhang, T.; Yu, L.; Qian, Z.; Liao, W.; Li, C. Degradation of the earthy and musty odorant 2,4,6-tricholoroanisole by persulfate activated with iron of different valences. Environ. Sci. Pollut. Res. 2017, 25, 3435–3445. [Google Scholar] [CrossRef]
  28. Aziz, K.H.H. Application of different advanced oxidation processes for the removal of chloroacetic acids using a planar falling film reactor. Chemosphere 2019, 228, 377–383. [Google Scholar] [CrossRef]
  29. Nam, S.-N.; Choong, C.E.; Hoque, S.; Farouk, T.I.; Cho, J.; Jang, M.; Snyder, S.A.; Meadows, M.E.; Yoon, Y. Catalytic non-thermal plasma treatment of endocrine disrupting compounds, pharmaceuticals, and personal care products in aqueous solution: A review. Chemosphere 2022, 290, 133395. [Google Scholar] [CrossRef]
  30. Gao, Y.; Wang, Q.; Ji, G.; Li, A. Degradation of antibiotic pollutants by persulfate activated with various carbon materials. Chem. Eng. J. 2021, 429, 132387. [Google Scholar] [CrossRef]
  31. Peng, W.; Dong, Y.; Fu, Y.; Wang, L.; Li, Q.; Liu, Y.; Fan, Q.; Wang, Z. Non-radical reactions in persulfate-based homogeneous degradation processes: A review. Chem. Eng. J. 2021, 421, 127818. [Google Scholar] [CrossRef]
  32. Ao, X.-W.; Eloranta, J.; Huang, C.-H.; Santoro, D.; Sun, W.-J.; Lu, Z.-D.; Li, C. Peracetic acid-based advanced oxidation processes for decontamination and disinfection of water: A review. Water Res. 2021, 188, 116479. [Google Scholar] [CrossRef] [PubMed]
  33. Collazo, C.; Charles, F.; Aguiló-Aguayo, I.; Marín-Sáez, J.; Lafarga, T.; Abadias, M.; Viñas, I. Decontamination of Listeria innocua from fresh-cut broccoli using UV-C applied in water or peroxyacetic acid, and dry-pulsed light. Innov. Food Sci. Emerg. Technol. 2019, 52, 438–449. [Google Scholar] [CrossRef] [Green Version]
  34. de Souza, J.B.; Valdez, F.Q.; Jeranoski, R.F.; Vidal, C.M.D.S.; Cavallini, G.S. Water and Wastewater Disinfection with Peracetic Acid and UV Radiation and Using Advanced Oxidative Process PAA/UV. Int. J. Photoenergy 2015, 2015, 860845. [Google Scholar] [CrossRef] [Green Version]
  35. Henao, L.D.; Turolla, A.; Antonelli, M. Disinfection by-products formation and ecotoxicological effects of effluents treated with peracetic acid: A review. Chemosphere 2018, 213, 25–40. [Google Scholar] [CrossRef]
  36. Hollman, J.; Dominic, J.A.; Achari, G. Degradation of pharmaceutical mixtures in aqueous solutions using UV/peracetic acid process: Kinetics, degradation pathways and comparison with UV/H2O2. Chemosphere 2020, 248, 125911. [Google Scholar] [CrossRef]
  37. Jahangiri-Rad, M.; Nadafi, K.; Mesdaghinia, A.; Nabizadeh, R.; Younesian, M.; Rafiee, M. Sequential study on reactive blue 29 dye removal from aqueous solution by peroxy acid and single wall carbon nanotubes: Experiment and theory. Iran. J. Environ. Health Sci. Eng. 2013, 10, 5. [Google Scholar] [CrossRef] [Green Version]
  38. Chhetri, R.K.; Baun, A.; Andersen, H.R. Acute toxicity and risk evaluation of the CSO disinfectants performic acid, peracetic acid, chlorine dioxide and their by-products hydrogen peroxide and chlorite. Sci. Total Environ. 2019, 677, 1–8. [Google Scholar] [CrossRef]
  39. Fraisse, A.; Temmam, S.; Deboosere, N.; Guillier, L.; Delobel, A.; Maris, P.; Vialette, M.; Morin, T.; Perelle, S. Comparison of chlorine and peroxyacetic-based disinfectant to inactivate Feline calicivirus, Murine norovirus and Hepatitis A virus on lettuce. Int. J. Food Microbiol. 2011, 151, 98–104. [Google Scholar] [CrossRef]
  40. Kiejza, D.; Kotowska, U.; Polińska, W.; Karpińska, J. Peracids—New oxidants in advanced oxidation processes: The use of peracetic acid, peroxymonosulfate, and persulfate salts in the removal of organic micropollutants of emerging concern—A review. Sci. Total Environ. 2021, 790, 148195. [Google Scholar] [CrossRef]
  41. Maurício, R.; Jorge, J.; Dias, R.; Noronha, J.P.; Amaral, L.; Daam, M.A.; Mano, A.P.; Diniz, M.S. The use of peracetic acid for estrogen removal from urban wastewaters: E2 as a case study. Environ. Monit. Assess. 2020, 192, 114. [Google Scholar] [CrossRef]
  42. Deng, J.; Wang, H.; Fu, Y.; Liu, Y. Phosphate-induced activation of peracetic acid for diclofenac degradation: Kinetics, influence factors and mechanism. Chemosphere 2022, 287, 132396. [Google Scholar] [CrossRef]
  43. Jian, L. Degradation of Antibiotics in Water by Advanced Oxidation Technologies Based on the Application of Peroxide. Master’s Thesis, Suzhou University of Science and Technology, Suzhou, China, 2019. [Google Scholar]
  44. Liu, B.; Guo, W.; Jia, W.; Wang, H.; Si, Q.; Zhao, Q.; Luo, H.; Jiang, J.; Ren, N. Novel Nonradical Oxidation of Sulfonamide Antibiotics with Co(II)-Doped g-C3N4-Activated Peracetic Acid: Role of High-Valent Cobalt–Oxo Species. Environ. Sci. Technol. 2021, 55, 12640–12651. [Google Scholar] [CrossRef]
  45. Shah, A.D.; Liu, Z.-Q.; Salhi, E.; Höfer, T.; von Gunten, U. Peracetic Acid Oxidation of Saline Waters in the Absence and Presence of H2O2: Secondary Oxidant and Disinfection Byproduct Formation. Environ. Sci. Technol. 2015, 49, 1698–1705. [Google Scholar] [CrossRef]
  46. Ghanbari, F.; Giannakis, S.; Lin, K.-Y.A.; Wu, J.; Madihi-Bidgoli, S. Acetaminophen degradation by a synergistic peracetic acid/UVC-LED/Fe(II) advanced oxidation process: Kinetic assessment, process feasibility and mechanistic considerations. Chemosphere 2021, 263, 128119. [Google Scholar] [CrossRef]
  47. Wang, J.; Xiong, B.; Miao, L.; Wang, S.; Xie, P.; Wang, Z.; Ma, J. Applying a novel advanced oxidation process of activated peracetic acid by CoFe2O4 to efficiently degrade sulfamethoxazole. Appl. Catal. B Environ. 2021, 280, 119422. [Google Scholar] [CrossRef]
  48. Zhou, F.; Lu, C.; Yao, Y.; Sun, L.; Gong, F.; Li, D.; Pei, K.; Lu, W.; Chen, W. Activated carbon fibers as an effective metal-free catalyst for peracetic acid activation: Implications for the removal of organic pollutants. Chem. Eng. J. 2015, 281, 953–960. [Google Scholar] [CrossRef]
  49. Da Silva, W.P.; Carlos, T.D.; Cavallini, G.S.; Pereira, D.H. Peracetic acid: Structural elucidation for applications in wastewater treatment. Water Res. 2020, 168, 115143. [Google Scholar] [CrossRef]
  50. Keller, B.; Wojcik, M.; Fletcher, T. A directly-dissociative stepwise reaction mechanism for gas-phase peroxyacetic acid. J. Photochem. Photobiol. A Chem. 2008, 195, 10–22. [Google Scholar] [CrossRef]
  51. Swern, D.; Silbert, L.S. Studies in the Structure of Organic Peroxides. Anal. Chem. 1963, 35, 880–885. [Google Scholar] [CrossRef]
  52. Rittenhouse, J.R.; Lobunez, W.; Swern, D.; Miller, J.G. The Electric Moments of Crganic Peroxides. II. Aliphatic Peracids1. J. Am. Chem. Soc. 1958, 80, 4850–4852. [Google Scholar] [CrossRef]
  53. Zhang, K.; Zhou, X.; Du, P.; Zhang, T.; Cai, M.; Sun, P.; Huang, C.-H. Oxidation of β-lactam antibiotics by peracetic acid: Reaction kinetics, product and pathway evaluation. Water Res. 2017, 123, 153–161. [Google Scholar] [CrossRef]
  54. Kim, J.; Zhang, T.; Liu, W.; Du, P.; Dobson, J.T.; Huang, C.-H. Advanced Oxidation Process with Peracetic Acid and Fe(II) for Contaminant Degradation. Environ. Sci. Technol. 2019, 53, 13312–13322. [Google Scholar] [CrossRef]
  55. Zhang, L.; Chen, J.; Zhang, Y.; Yu, Z.; Ji, R.; Zhou, X. Activation of peracetic acid with cobalt anchored on 2D sandwich-like MXenes (Co@MXenes) for organic contaminant degradation: High efficiency and contribution of acetylperoxyl radicals. Appl. Catal. B Environ. 2021, 297, 120475. [Google Scholar] [CrossRef]
  56. Wang, Z.; Wang, J.; Xiong, B.; Bai, F.; Wang, S.; Wan, Y.; Zhang, L.; Xie, P.; Wiesner, M.R. Application of Cobalt/Peracetic Acid to Degrade Sulfamethoxazole at Neutral Condition: Efficiency and Mechanisms. Environ. Sci. Technol. 2020, 54, 464–475. [Google Scholar] [CrossRef]
  57. Cao, C.; Zhang, S.; Zhang, F.; Zhang, K. Research Progress of Peracetic Acid (PAA): An Emerging Disinfectant in Drinking Water. China Water Wastewater 2018, 34, 36–40. [Google Scholar] [CrossRef]
  58. Ison, S.; Beattie, M. Disinfection, sterilization and preservation (5th ed). Aust. Infect. Control 2002, 7, 74. [Google Scholar] [CrossRef]
  59. Kitis, M. Disinfection of wastewater with peracetic acid: A review. Environ. Int. 2004, 30, 47–55. [Google Scholar] [CrossRef]
  60. Pohjanvesi, S.; Mustonen, E.L.; Pukkinen, A.; Lehtinen, R. Process for the Production of Percarboxylic Acid. U.S. Patent 177732[P], 4 April 2002. [Google Scholar]
  61. Palani, A.; Pandurangan, A. Single pot synthesis of peroxyacetic acid from acetic acid and hydrogen peroxide using various solid acid catalysts. Catal. Commun. 2006, 7, 875–878. [Google Scholar] [CrossRef]
  62. Saha, M.S.; Denggerile, A.; Nishiki, Y.; Furuta, T.; Ohsaka, T. Synthesis of peroxyacetic acid using in situ electrogenerated hydrogen peroxide on gas diffusion electrode. Electrochem. Commun. 2003, 5, 445–448. [Google Scholar] [CrossRef]
  63. Cai, M.; Sun, P.; Zhang, L.; Huang, C.-H. UV/Peracetic Acid for Degradation of Pharmaceuticals and Reactive Species Evaluation. Environ. Sci. Technol. 2017, 51, 14217–14224. [Google Scholar] [CrossRef]
  64. Chen, S.; Cai, M.; Liu, Y.; Zhang, L.; Feng, L. Effects of water matrices on the degradation of naproxen by reactive radicals in the UV/peracetic acid process. Water Res. 2019, 150, 153–161. [Google Scholar] [CrossRef]
  65. Davies, D.M.; Deary, M.E. Determination of peracids in the presence of a large excess of hydrogen peroxide using a rapid and convenient spectrophotometric method. Analyst 1988, 113, 1477–1479. [Google Scholar] [CrossRef]
  66. Tashkhourian, J.; Hormozi-Nezhad, M.R.; Khodaveisi, J.; Dashti, R. Localized surface plasmon resonance sensor for simultaneous kinetic determination of peroxyacetic acid and hydrogen peroxide. Anal. Chim. Acta 2013, 762, 87–93. [Google Scholar] [CrossRef]
  67. Zhang, T.; Huang, C.-H. Modeling the Kinetics of UV/Peracetic Acid Advanced Oxidation Process. Environ. Sci. Technol. 2020, 54, 7579–7590. [Google Scholar] [CrossRef]
  68. Zhao, Z.; Li, X.; Li, H.; Qian, J.; Pan, B. New Insights into the Activation of Peracetic Acid by Co(II): Role of Co(II)-Peracetic Acid Complex as the Dominant Intermediate Oxidant. ACS ES T Eng. 2021, 1, 1432–1440. [Google Scholar] [CrossRef]
  69. Manoli, K.; Li, R.; Kim, J.; Feng, M.; Huang, C.-H.; Sharma, V.K. Ferrate(VI)-peracetic acid oxidation process: Rapid degradation of pharmaceuticals in water. Chem. Eng. J. 2022, 429, 132384. [Google Scholar] [CrossRef]
  70. Carlos, T.D.; Bezerra, L.B.; Vieira, M.M.; Sarmento, R.A.; Pereira, D.H.; Cavallini, G.S. Fenton-type process using peracetic acid: Efficiency, reaction elucidations and ecotoxicity. J. Hazard. Mater. 2021, 403, 123949. [Google Scholar] [CrossRef]
  71. Chen, S. Degradation Mechanisms of Nap and Effects of Water Matrices on Free Radicals Conversion in UV/Peracetic Acid Process. Master’s Thesis, Beijing Forestry University, Beijing, China, 2019. [Google Scholar]
  72. Luukkonen, T.; Heyninck, T.; Rämö, J.; Lassi, U. Comparison of organic peracids in wastewater treatment: Disinfection, oxidation and corrosion. Water Res. 2015, 85, 275–285. [Google Scholar] [CrossRef]
  73. Maurício, R.; Semedo, F.; Dias, R.; Noronha, J.P.; Amaral, L.; Daam, M.; Mano, A.P.; Diniz, M. Efficacy assessment of peracetic acid in the removal of synthetic 17α-ethinyl estradiol contraceptive hormone in wastewater. J. Environ. Sci. 2020, 89, 1–8. [Google Scholar] [CrossRef]
  74. Wu, J.; Zheng, X.; Wang, Y.; Liu, H.; Wu, Y.; Jin, X.; Chen, P.; Lv, W.; Liu, G. Activation of peracetic acid via Co3O4 with double-layered hollow structures for the highly efficient removal of sulfonamides: Kinetics insights and assessment of practical applications. J. Hazard. Mater. 2022, 431, 128579. [Google Scholar] [CrossRef]
  75. Wu, W.; Tian, D.; Liu, T.; Chen, J.; Huang, T.; Zhou, X.; Zhang, Y. Degradation of organic compounds by peracetic acid activated with Co3O4: A novel advanced oxidation process and organic radical contribution. Chem. Eng. J. 2020, 394, 124938. [Google Scholar] [CrossRef]
  76. Hey, G.; Ledin, A.; Jansen, J.L.C.; Andersen, H.R. Removal of pharmaceuticals in biologically treated wastewater by chlorine dioxide or peracetic acid. Environ. Technol. 2012, 33, 1041–1047. [Google Scholar] [CrossRef]
  77. Kim, J.; Du, P.; Liu, W.; Luo, C.; Zhao, H.; Huang, C.-H. Cobalt/Peracetic Acid: Advanced Oxidation of Aromatic Organic Compounds by Acetylperoxyl Radicals. Environ. Sci. Technol. 2020, 54, 5268–5278. [Google Scholar] [CrossRef]
  78. Wang, H.; Deng, J.; Lu, X.; Wan, L.; Huang, J.; Liu, Y. Rapid and continuous degradation of diclofenac by Fe(II)-activated persulfate combined with bisulfite. Sep. Purif. Technol. 2021, 262, 118335. [Google Scholar] [CrossRef]
  79. Zhou, X.; Luo, H.; Sheng, B.; Chen, X.; Wang, Y.; Chen, Q.; Zhou, J. Cu2+/Cu+ cycle promoted PMS decomposition with the assistance of Mo for the degradation of organic pollutant. J. Hazard. Mater. 2021, 411, 125050. [Google Scholar] [CrossRef]
  80. Wang, Z.; Fu, Y.; Peng, Y.; Wang, S.; Liu, Y. HCO3/CO32– enhanced degradation of diclofenac by Cu(II)-activated peracetic acid: Efficiency and mechanism. Sep. Purif. Technol. 2021, 277, 119434. [Google Scholar] [CrossRef]
  81. Zhang, L.; Fu, Y.; Wang, Z.; Zhou, G.; Zhou, R.; Liu, Y. Removal of diclofenac in water using peracetic acid activated by zero valent copper. Sep. Purif. Technol. 2021, 276, 119319. [Google Scholar] [CrossRef]
  82. Jingxiao, W.; Kean, Z.; Fei, C. Degradation performance and mechanism of Rhodamine B by chloride Ion activated peracetic acid. Res. Environ. Sci. 2021, 34, 2850–2858. [Google Scholar] [CrossRef]
  83. Haysar, A.; Jia, J.; Ren, D.; Liao, Y.; Wang, J.; Zhao, M. The removal of hpam with fenton-like reagents oxidation process. Spec. Petrochem. 2016, 33, 4. [Google Scholar] [CrossRef]
  84. Wang, Z.; Shi, H.; Wang, S.; Liu, Y.; Fu, Y. Degradation of diclofenac by Fe(II)-activated peracetic acid. Environ. Technol. 2021, 42, 4333–4341. [Google Scholar] [CrossRef] [PubMed]
  85. Tian, D.; Wu, W.; Shen, Z.; Huang, T.; Chen, J. Degradation of organic dyes with peracetic acid activated by Co(II). Acta Sci. Circumstantiae 2018, 38, 4023–4031. [Google Scholar] [CrossRef]
  86. Rothbart, S.; Ember, E.E.; van Eldik, R. Mechanistic studies on the oxidative degradation of Orange II by peracetic acid catalyzed by simple manganese(ii) salts. Tuning the lifetime of the catalyst. New J. Chem. 2012, 36, 732–748. [Google Scholar] [CrossRef]
  87. Yuan, D.; Yang, K.; Pan, S.; Xiang, Y.; Tang, S.; Huang, L.; Sun, M.; Zhang, X.; Jiao, T.; Zhang, Q.; et al. Peracetic acid enhanced electrochemical advanced oxidation for organic pollutant elimination. Sep. Purif. Technol. 2021, 276, 119317. [Google Scholar] [CrossRef]
  88. Bezerra, L.B.; Carlos, T.D.; das Neves, A.P.N.; Durães, W.A.; Sarmento, R.D.A.; Pereira, D.H.; Cavallini, G.S. Theoretical-experimental study of the advanced oxidative process using peracetic acid and solar radiation: Removal efficiency and thermodynamic elucidation of radical formation processes. J. Photochem. Photobiol. A Chem. 2022, 423, 113615. [Google Scholar] [CrossRef]
  89. Wang, J.; Wan, Y.; Ding, J.; Wang, Z.; Ma, J.; Xie, P.; Wiesner, M.R. Thermal Activation of Peracetic Acid in Aquatic Solution: The Mechanism and Application to Degrade Sulfamethoxazole. Environ. Sci. Technol. 2020, 54, 14635–14645. [Google Scholar] [CrossRef]
  90. Daswat, D.P.; Mukhopadhyay, M. Photochemical degradation of chlorophenol industry wastewater using peroxy acetic acid (PAA). Chem. Eng. J. 2012, 209, 1–6. [Google Scholar] [CrossRef]
  91. Rizzo, L.; Lofrano, G.; Gago, C.; Bredneva, T.; Iannece, P.; Pazos, M.; Krasnogorskaya, N.; Carotenuto, M. Antibiotic contaminated water treated by photo driven advanced oxidation processes: Ultraviolet/H2O2 vs ultraviolet/peracetic acid. J. Clean. Prod. 2018, 205, 67–75. [Google Scholar] [CrossRef]
  92. Li, Z.; Yong-sheng, F.; Yi-qing, L. Degradation of diclofenac in water by Cu2+ enhanced UV activation of peracetic acid. China Environ. Sci. 2020, 40, 5260–5269. [Google Scholar] [CrossRef]
  93. Yinhao, D.; Shaogui, Y.; Chengdu, Q.; Shiyin, L.; Xiaolong, D.; Huan, H.; Haiou, S. Activation of peracetic acid process for aquatic organic pollutants degradation: A review. Environ. Chem. Lett. 2021, 40, 497–508. [Google Scholar] [CrossRef]
  94. Tang, L.; Long, K.; Chen, S.; Gui, D.; He, C.; Li, J.; Tao, X. Removal of thiophene sulfur model compound for coal by microwave with peroxyacetic acid. Fuel 2020, 272, 117748. [Google Scholar] [CrossRef]
  95. Chin, J.F.; Heng, Z.W.; Teoh, H.C.; Chong, W.C.; Pang, Y.L. Recent development of magnetic biochar crosslinked chitosan on heavy metal removal from wastewater—Modification, application and mechanism. Chemosphere 2022, 291, 133035. [Google Scholar] [CrossRef]
  96. Yang, F.; Zhang, S.; Sun, Y.; Du, Q.; Song, J.; Tsang, D.C. A novel electrochemical modification combined with one-step pyrolysis for preparation of sustainable thorn-like iron-based biochar composites. Bioresour. Technol. 2019, 274, 379–385. [Google Scholar] [CrossRef]
  97. Yang, J.; Zhao, Y.; Ma, S.; Zhu, B.; Zhang, J.; Zheng, C. Mercury Removal by Magnetic Biochar Derived from Simultaneous Activation and Magnetization of Sawdust. Environ. Sci. Technol. 2016, 50, 12040–12047. [Google Scholar] [CrossRef]
  98. Wang, S.; Wang, H.; Liu, Y.; Fu, Y. Effective degradation of sulfamethoxazole with Fe2+-zeolite/peracetic acid. Sep. Purif. Technol. 2020, 233, 115973. [Google Scholar] [CrossRef]
  99. Rokhina, E.V.; Makarova, K.; Lahtinen, M.; Golovina, E.A.; Van As, H.; Virkutyte, J. Ultrasound-assisted MnO2 catalyzed homolysis of peracetic acid for phenol degradation: The assessment of process chemistry and kinetics. Chem. Eng. J. 2013, 221, 476–486. [Google Scholar] [CrossRef]
  100. Rokhina, E.V.; Makarova, K.; Golovina, E.A.; Van As, H.; Virkutyte, J. Free Radical Reaction Pathway, Thermochemistry of Peracetic Acid Homolysis, and Its Application for Phenol Degradation: Spectroscopic Study and Quantum Chemistry Calculations. Environ. Sci. Technol. 2010, 44, 6815–6821. [Google Scholar] [CrossRef]
  101. 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]
  102. Wang, J.; Wang, Z.; Cheng, Y.; Cao, L.; Bai, F.; Yue, S.; Xie, P.; Ma, J. Molybdenum disulfide (MoS2): A novel activator of peracetic acid for the degradation of sulfonamide antibiotics. Water Res. 2021, 201, 117291. [Google Scholar] [CrossRef]
  103. Huang, T.; Chen, J.; Shen, Z.; Tian, D. A Method for Degrading Azo Dye Golden Orange G in Wastewater. CN108298668A, 20 July 2018. [Google Scholar]
  104. Zhang, L.; Liu, Y.; Fu, Y. Degradation kinetics and mechanism of diclofenac by UV/peracetic acid. RSC Adv. 2020, 10, 9907–9916. [Google Scholar] [CrossRef] [Green Version]
  105. Mukhopadhyay, M.; Daswat, D.P. Kinetic and mechanistic study of photochemical degradation of 4-chlorophenol using peroxy acetic acid (PAA). Desalination Water Treat. 2014, 52, 5704–5714. [Google Scholar] [CrossRef]
  106. Izadifard, M.; Achari, G.; Langford, C.H. Degradation of sulfolane using activated persulfate with UV and UV-Ozone. Water Res. 2017, 125, 325–331. [Google Scholar] [CrossRef] [PubMed]
Figure 1. (a) PAA molecular formula; Molecular electrostatic potential diagram of (b) PAA and (c) H2O2.
Figure 1. (a) PAA molecular formula; Molecular electrostatic potential diagram of (b) PAA and (c) H2O2.
Water 14 02309 g001
Figure 2. UV/PAA reaction scheme. Reprinted with permission from Ref. [67]. 2020, Tianqi Zhang, Ching-Hua Huang.
Figure 2. UV/PAA reaction scheme. Reprinted with permission from Ref. [67]. 2020, Tianqi Zhang, Ching-Hua Huang.
Water 14 02309 g002
Figure 3. Transition metal/PAA reaction scheme.
Figure 3. Transition metal/PAA reaction scheme.
Water 14 02309 g003
Figure 4. Mechanism of activated carbon catalyzed PAA. Reprinted with permission from Ref. [48]. 2015, Fengya Zhou, Chao Lu, Yuyuan Yao, Lijie Sun, Fei Gong, Daiwen Li, Kemei Pei, Wangyang Lu, Wenxing Chen.
Figure 4. Mechanism of activated carbon catalyzed PAA. Reprinted with permission from Ref. [48]. 2015, Fengya Zhou, Chao Lu, Yuyuan Yao, Lijie Sun, Fei Gong, Daiwen Li, Kemei Pei, Wangyang Lu, Wenxing Chen.
Water 14 02309 g004
Figure 5. Non-radical pathway of Co (II)-PAA. Reprinted with permission from Ref. [68]. 2021, Zihao Zhao, Xinhong Li, Hongchao Li, Jieshu Qian, Bingcai Pan. (a), Co (II)-doped g-C3N4-PAA. Reprinted with permission from Ref. [44]. 2021, Banghai Liu, Wanqian Guo, Wenrui Jia, Huazhe Wang, Qishi Si, Qi Zhao, Haichao Luo, Jin Jiang, and Nanqi Ren. (b) process.
Figure 5. Non-radical pathway of Co (II)-PAA. Reprinted with permission from Ref. [68]. 2021, Zihao Zhao, Xinhong Li, Hongchao Li, Jieshu Qian, Bingcai Pan. (a), Co (II)-doped g-C3N4-PAA. Reprinted with permission from Ref. [44]. 2021, Banghai Liu, Wanqian Guo, Wenrui Jia, Huazhe Wang, Qishi Si, Qi Zhao, Haichao Luo, Jin Jiang, and Nanqi Ren. (b) process.
Water 14 02309 g005
Table 1. Properties of PAA and other oxides.
Table 1. Properties of PAA and other oxides.
ItemsValue
pKa8.2
Eh0(V)PAA1.0~1.96
H2O21.78
Cl1.48
ClO21.28
FeVI0.9~1.9
O32.08
Molar mass (g/mol)76.05
Density (g/mol)1.0375
boiling point (°C)110
melting point (°C)0.2
Henry’s law constant (M/atm)4.68 × 102
log Kow (at pH 7)−0.66
O-OH bond(kJ/mol)PAA159
H2O2213
Table 2. An overview of research on the direct oxidation of aqueous pollutants by PAA.
Table 2. An overview of research on the direct oxidation of aqueous pollutants by PAA.
ContaminantsC0 (μg·L−1)CPAA (mg·L−1)pHTime (min)Max. Removal Efficiency (%)MatrixesReferences
17 β-Estradiol50157.910–20100WW 1[41]
17 α-ethinyl estradiol50157.910–20100WW[73]
Acetaminophen20 mg·L−14 mM530~15DI water 2[46]
Reactive Brilliant Red X-3B50 μM5 mM740<5DI water[48]
sulfamethazine8 mg·L−10.4 mM6.360~20DI water[74]
2,4-dichlorophenol20 μM0.26 mM720<5DI water[55]
sulfamethazine10 μM100 μM7 ± 0.215<5DI water[56]
Clofibric acid0.05 mM0.5 mM790<5DI water[75]
Ibuprofen40506.7 (WW 1 3), 7.0 (WW 2 4 and WW 3 5)n.r. 6~45 (WW 1), ~20 (WW 2), <5 (WW 3)WW 1, WW 2, WW 3[76]
Naproxen~80 (WW 1), ~60 (WW 2), ~40 (WW 3)
Diclofenac~75 (WW 1), >95 (WW 2), ~20 (WW 3)
Mefenamic acid>95 (WW 1), ~90 (WW 2), ~90 (WW 3)
Gemfibrozil~75 (WW 1), ~20 (WW 2), ~20 (WW 3)
Clofibric acid~60 (WW 1), ~45 (WW 2), <10 (WW 3)
Blue 29 dye30 mg ·L−1CH3C(O)OH/dye = 344/13–860–18020.2~56.4DI water[37]
Note(s): 1 Wastewater. 2 Deionized water. 3 low COD WW. 4 medium COD WW. 5 high COD WW. 6 not reported.
Table 5. Studies of activating PAA in heterogeneous phase systems.
Table 5. Studies of activating PAA in heterogeneous phase systems.
CatalystsCCatalyst (g·L−1)ContaminantsC0CPAA (mM)pHTime
(min)
Max. Removal Efficiency (%)References
Ultrasound/MnO21Phenol98.8 mg·L−110072089[99]
MnO20.7Phenoln.r.50 mg·L−19.5120n.r.[100]
Co3O40.2Orange Gn.r.1790100[75]
LaCoO30.1SMX8 mg·L−10.4n.r.60100[74]
CoFe2O30.2SMXn.r.0.273074.7[47]
ACF2Reactive Brilliant Red X-3B30.8 mg·L−1380 mg·L−174092.5[48]
Co(II)/g-C3N40.1SMX10 μM0.24.2430~100[44]
Fe2+-zeolite0.8SMX5 μM400 μM750~100[101]
Co@MXenes10 mg·L−12,4-dichloropheno20 μM0.26720~100[55]
MoS20.3SMX10 μM0.3 31576.1[102]
Phenol-AC0.1–0.2Orange Gn.r.2.6 × 10−4~3.9 × 10−37n.r.Effective degradation[103]
N-rGO0.75SMX0.15 mM136096[43]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Shi, C.; Li, C.; Wang, Y.; Guo, J.; Barry, S.; Zhang, Y.; Marmier, N. Review of Advanced Oxidation Processes Based on Peracetic Acid for Organic Pollutants. Water 2022, 14, 2309. https://doi.org/10.3390/w14152309

AMA Style

Shi C, Li C, Wang Y, Guo J, Barry S, Zhang Y, Marmier N. Review of Advanced Oxidation Processes Based on Peracetic Acid for Organic Pollutants. Water. 2022; 14(15):2309. https://doi.org/10.3390/w14152309

Chicago/Turabian Style

Shi, Changjie, Cong Li, Yong Wang, Jiaqi Guo, Sadou Barry, Yunshu Zhang, and Nicolas Marmier. 2022. "Review of Advanced Oxidation Processes Based on Peracetic Acid for Organic Pollutants" Water 14, no. 15: 2309. https://doi.org/10.3390/w14152309

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