Next Article in Journal
Removal from Water of Some Pharmaceuticals by Photolysis and Photocatalysis: Kinetic Models
Previous Article in Journal
Graphene Oxide-Supported Metal Catalysts for Selective Hydrogenation of Cinnamaldehyde: Impact of Metal Choice and Support Structure
Previous Article in Special Issue
Bacterial Inactivation and Organic Pollutant Degradation in Slaughterhouse Wastewater Using Ag2O/Ba/TiO2 Nanocomposite
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Current Progress in Advanced Oxidation Processes for the Removal of Contaminants of Emerging Concern Using Peracetic Acid as an Effective Oxidant

by
Bakhta Bouzayani
1,2,
Sourour Chaâbane Elaoud
2 and
Maria Ángeles Sanromán
1,*
1
BIOSUV Group, Department of Chemical Engineering, Campus As Lagoas-Marcosende, CINTECX, University of Vigo, 36310 Vigo, Spain
2
Laboratory of Physical Chemistry of the Solid State, Department of Chemical, University of Sfax, Sfax 3000, Tunisia
*
Author to whom correspondence should be addressed.
Catalysts 2025, 15(5), 469; https://doi.org/10.3390/catal15050469 (registering DOI)
Submission received: 24 March 2025 / Revised: 29 April 2025 / Accepted: 8 May 2025 / Published: 10 May 2025

Abstract

:
The growing diversity and prevalence of contaminants of emerging concern (CECs) in aquatic environments present significant risks to human health and ecosystems, necessitating the development of effective remediation strategies. Advanced oxidation processes (AOPs) have emerged as a promising solution due to their ability to produce highly reactive species that efficiently degrade persistent contaminants. Among the various oxidizing agents, peracetic acid (PAA) has attracted significant attention in the field of water treatment for its powerful oxidative properties, environmentally safe decomposition, and ease of use. This article is designed to offer a comprehensive overview of the latest trends in PAA-based AOPs. The discussion begins with an overview of the intrinsic performance of PAA, emphasizing its oxidation potential and degradation mechanisms. Subsequently, the effectiveness of PAA-based AOPs in remediating CECs is explored, focusing on transition metal-mediated activation (Fe, Co, Mn), UV irradiation, and carbon-based catalysts, all of which enhance the generation of reactive species (RS). Next, the determination of RS in PAA-based AOPs is examined, distinguishing between free radical (organic and inorganic) and non-radical (singlet oxygen and high-valent metal) mechanisms that govern pollutant degradation. Then, key factors affecting the removal of CECs in PAA-based AOPs, including initial PAA concentration, catalyst dosage, and pH, are also addressed. Following that, the potential by-products and hazard assessments associated with PAA oxidation are discussed. Finally, current challenges and future research directions are proposed to facilitate the large-scale application of PAA-based AOPs in water remediation.

Graphical Abstract

1. Introduction

The quality and availability of aquatic resources are crucial for sustaining the health of all organisms within ecosystems, ranging from microscopic species to humans [1,2,3]. Access to clean water supports ecosystem stability and enhances human health and well-being. However, in recent times, an increasing presence of unconventional organic pollutants has been detected in natural water bodies, many of which lack regulatory standards, posing significant threats to both ecosystem stability and public health even at environmentally relevant concentrations [4]. The Environmental Protection Agency refers to these pollutants as emerging contaminants; however, since the mid-1990s, the term ‘contaminants of emerging concern’ (CECs) has been commonly used to describe substances such as pesticides, personal care products, industrial chemicals, pharmaceuticals, veterinary antibiotics, and nanomaterials. CECs pose significant risks to human health and aquatic organisms even at low concentrations [5] due to their persistence and bioaccumulation, which is attributed to their high resistance to degradation, a quality linked to their complex structure [6]. Consequently, CEC exposure in humans has been linked to a range of adverse health outcomes, including reproductive system disorders, thyroid dysfunction, Alzheimer’s disease, cancer, and obesity [7]. In wildlife, CECs can also impair reproductive systems, reduce hatchability, and alter vitellogenin levels [7].
Nevertheless, conventional wastewater treatment strategies, including physicochemical and biological methods, often exhibit limited efficiency and may fail to meet the stringent standards imposed by contemporary environmental regulations [8]. Physical methods primarily remove contaminants from water, which may not be adequate for the complete mineralization of pollutants [9]. On the other hand, biological methods may not be suitable for urgent remediation, as they require prolonged periods of use for effective treatment. Furthermore, they are sensitive to changes in pH, temperature, and aeration, all of which are crucial for microbes to efficiently degrade pollutants [10]. These conventional pollutant treatment methods, still widely used for removing most contaminants, exhibit limited efficiency, as they fail to achieve complete destruction. As a result, there is a growing interest in sustainable processes that can better address the challenges posed by these contaminants.
Advanced oxidation processes (AOPs) have emerged as a promising solution for contaminant removal, owing to their environmentally friendly characteristics and wide applicability [11,12]. These processes bridge the gap between traditional physicochemical and biological methods and the current restrictions imposed by environmental legislation [9]. Through the generation of potent reactive radicals, such as sulfate radicals ( S O 4 ) [13,14,15], hydroxyl radicals ( H O ) [16,17,18], and peroxyl radicals ( R O O ) [19,20,21], AOPs effectively promote the breakdown of bulky organic pollutants into more benign molecules, ultimately leading to their complete mineralization into C O 2 and H 2 O [22].
The generation of reactive intermediates is initiated by the activation of oxidants, with the nature of the radicals produced determined by the specific type of oxidant used. In AOPs, commonly used oxidants include hydrogen peroxide H 2 O 2 [23], ozone O 3 [24], persulfate S 2 O 8 2 [25], and peroxymonosulfate (PMS) [26]. Since 2019, PAA ( C H 3 C O O O H ) has gained significant attention as a promising candidate due to its potent oxidative potential, durability, high efficiency at ambient temperature, widespread availability, lower by-product formation, and reasonable cost [27,28,29].
Additionally, its lower O-O bond energy (38 kcal·mol−1) compared to PMS (75.7 kcal·mol−1) and H 2 O 2 (51 kcal·mol−1) indicates that PAA requires less energy for activation, further enhancing its appeal as an oxidant [30]. This reduced energy requirement facilitates the generation of reactive species (RS) in PAA-based AOPs, which are known to produce H O and various organic radicals, including acetylperoxyl radicals   C H 3 C O O O , acetyloxy radicals C H 3 C ( O ) O , and alkoxyl radicals ( R O ) [19]. Notably, ( R O ) radicals are unique to PAA-based AOPs, imparting distinctive characteristics that set them apart from conventional AOPs. Furthermore, PAA not only generates free radicals but also promotes the formation of non-radical species such as singlet oxygen (1 O 2 ) [31], showcasing its versatile oxidizing potential. This versatility positions PAA as a crucial component in advanced water treatment technologies, especially considering the urgent need to manage CECs.
The benefits of PAA make it a compelling candidate for alternative, efficient approaches to removing CECs. In addition, as illustrated in Figure 1, the combination of UV light and PAA can be utilized to achieve simultaneous pathogen inactivation and micropollutant degradation. Furthermore, significant attention has been devoted to understanding the microbial inactivation kinetics of PAA and evaluating how key experimental parameters influence the overall process efficiency.
Therefore, this review aims to explore the role of PAA in AOPs, focusing on its intrinsic performance and activation methods, including metal catalysts, UV irradiation, and carbon-based catalysts. Additionally, we discuss its efficiency in degrading CECs and the determination of the reactive species involved. Furthermore, we examine the various variables influencing its performance. Lastly, we highlight the challenges associated with PAA-based processes and propose potential future avenues for enhancing their applicability in environmental remediation.

2. Intrinsic Performance of PAA

Due to its strong microbicidal effects, PAA has been widely utilized [32]. It is a member of the organic peracid family, characterized by the presence of a -COOOH group. Due to its high redox potential, which ranges from 1.06 to 1.96 V in alkaline and acidic conditions, respectively, PAA has the capability to break down organic micropollutants in water [19,32]. Furthermore, the peroxide group of PAA is thermodynamically unstable. This instability arises from electron repulsion between the oxygen atoms in the peroxide moiety, which further weakens the O-O bond and increases its tendency to decompose, forming hydroxyl and aliphatic radicals. These RS facilitate the degradation of organic pollutants; for this reason, PAA is regarded as an oxidizing agent.
A growing number of recent studies emphasize the use of PAA for treating water contaminated with CECs, underscoring its effectiveness in addressing these challenges. This potency is owed to the distinctive properties of PAA, which functions as a moderately specific oxidant. For instance, Zhang et al. [33] indicated that PAA can directly react with contaminants containing electron-rich, dense moieties, such as unsaturated bonds or conjugated systems. Owing to their high electron density, these groups exhibit greater reactivity toward oxidation, thereby increasing the likelihood of interacting with PAA. As a result, PAA has been shown to effectively oxidize several β-lactam antibiotics, recognized as CECs, including cefalexin, cefadroxil, cefapirin, cephalothin, ampicillin, amoxicillin, and penicillin [33]. Their reactivity with PAA is attributed to their nucleophilic nature, as evidenced by the correlation between the reaction rate constant and the global nucleophilicity of these compounds. Additionally, contaminants containing sulfur moieties exhibit exceptionally high reactivity toward PAA, as the presence of electron-donating groups enhances their interaction with electrophilic species [34]. To improve the understanding of reactions triggered by PAA, Kim et al. [35] published a comprehensive review on the behavior of PAA with the organic compounds of diverse structures and explored the underlying mechanisms. Demonstrating its selectivity, the second-order rate constants for the PAA oxidation of these compounds span nearly 10 orders of magnitude, ranging from 3.2 × 10 6   M 1 · s 1 to 1 × 10 5   M 1 · s 1 .
However, many organic pollutants, such as ibuprofen, gemfibrozil, and naproxen (NAP), are resistant to degradation at low PAA doses, as PAA is unable to effectively break them down [36]. This limitation is primarily attributed to its low homolytic cleavage rate of 6 × 10 12   s 1 , which restricts its effectiveness as an oxidant in technical processes [32]. Moreover, PAA’s high selectivity prevents it from effectively degrading recalcitrant pollutants on its own [36]. This underscores the need for activators to efficiently produce potent oxidizing agents.

3. Proficiency of PAA-Based AOPs in CECs Remediation

In AOPs, PAA has garnered attention as an oxidizing agent capable of producing peroxyl radicals ( C H 3 C O O O , C H 3 C ( O ) O ) along with H O radicals through its activation [37]. The formation of free radicals in PAA-based AOPs was first reported by Caretti and Lubello, who implemented a UV/PAA approach in a wastewater disinfection pilot plant [38]. However, only in recent years have PAA-based AOPs gained recognition as a method for eliminating CECs. Relative to H 2 O 2 , which is considered the primary oxidant in AOPs, the lower unoccupied molecular orbital (LUMO) energy of PAA (−0.25 eV) indicates its easier activation, suggesting that PAA has a greater capacity to accept electrons than H 2 O 2 (LUMO energy = 0.57 eV) [39]. Therefore, PAA-based AOPs are regarded as a promising advancement over conventional H 2 O 2 -based AOPs, with a growing interest in these processes since 2019.
Various methods have been proposed to activate PAA through the cleavage of the O-O bond into two radicals, H O and C H 3 C ( O ) O (Equation (1)), an important step in the oxidation of micropollutants. Among these, Zhang and Huang proposed a widely accepted mechanism describing the reactions of different radicals in a UV/PAA system, which has since been applied to other PAA-based processes [40]. In addition to degrading organic pollutants, both radicals are also capable of interacting with PAA to produce C H 3 C O O O (Equations (2) and (3)). These C H 3 C O O O radicals can, in turn, undergo bimolecular decomposition, yielding O 2 and C H 3 C ( O ) O (Equation (4)). Moreover, the acetoxy radicals formed are also capable of undergoing radical–radical coupling (Equation (5)). On the other hand, to produce methyl radicals ( C H 3 ), C H 3 C ( O ) O undergoes decarboxylation (Equation (6)). Due to their instability, C H 3   are reactive intermediates that react immediately upon contact with oxygen to generate peroxyl radicals (Equation (7)), which subsequently undergo bimolecular breakdown (Equations (8)–(10)).
C H 3 C O O O H     C H 3 C ( O ) O + H O
C H 3 C O O O H   + H O     C H 3 C O O O   + H 2 O  
C H 3 C O O O H   + C H 3 C ( O ) O     C H 3 C O O O + C H 3 C O O H
C H 3 C O O O   + C H 3 C O O   O     2   C H 3 C ( O ) O   + O 2
C H 3 C ( O ) O   + C H 3 C ( O ) O     ( C H 3 C ( O ) O ) 2
C H 3 C O O     C H 3 + C O 2
C H 3 + O 2     C H 3 O O
C H 3 O O   + C H 3 O O     H C H O   + C H 3 O H   + O 2
C H 3 O O   +   C H 3 O O     2 H C H O   + H 2 O 2
C H 3 O O   + C H 3 O O     2 C H 3 O + O 2
To initiate another oxidation cycle, H O reacts with C H 3 C ( O ) O again to generate a new PAA molecule, as shown in Equation (11) [40].
C H 3 C ( O ) O   + H O     C H 3 C O O O H
Diverse technologies have been applied for PAA activation to facilitate the formation of RS (Figure 1), including the following:
-
Transition metal ions, such as iron [39], cobalt [20,41,42], manganese [43], copper [44], ruthenium [45], silver [46], chromium [47], and molybdenum [48,49], which enhance reactions with high efficiency.
-
Heterogeneous metal catalysts offering stability and reusability [50,51,52].
-
UV light irradiation (UV/PAA) for increased radical production [21,53,54].
-
Several energy sources, such as heat [55], ultrasound [56], electricity [57], and microwaves [58], to trigger activation under different conditions.
-
Non-metallic carbon catalysts, presenting a green and sustainable alternative [59,60].

3.1. Transition Metal for the Activation of PAA

AOPs that involve the activation of PAA by transition metal ions seem highly promising [61]. When transition metals are introduced into a PAA solution, they can easily decompose the O-O bond, leading to the production of C H 3 C O O O and C H 3 C ( O ) O , as evidenced by the following (Equations (12) and (13)). Although PAA activation by transition metal catalysts closely mirrors the traditional Fenton process, it demonstrates greater efficiency in radical generation [62]. This enhanced efficiency is reflected in the reactivity of PAA, which, when combined with F e 2 + , is more than 650 times greater than that of H 2 O 2 [39]. Additionally, the reaction rate constant for radical generation from PAA is significantly higher (ranging from 0.5 to 1.1 × 10 5   M 1 · s 1 ) within the pH range of 3 to 7.1, compared to the much lower rate observed for the Fenton reaction involving F e 2 + / H 2 O 2 (K = 63 to 76 M 1 · s 1 ) [39]. This comparison highlights the considerably enhanced reactivity and radical generation efficiency of PAA activation relative to the traditional Fenton process, primarily because PAA decomposition is the main source of H O radicals in Fe(II)/PAA systems.
M n + + C H 3 C O O O H     M n + 1 + + C H 3 C ( O ) O + H O
M ( n + 1 ) +   + C H 3 C O O O H     M n + + C H 3 C O O O + H +
Several recent works have investigated the use of transition metals for PAA activation. While the majority of this research has focused on dissolved metal catalysts, some have also explored the use of heterogeneous metal catalysts, such as zero-valent metals and metal oxides. A summary of these findings is presented in Table 1.

3.1.1. Iron-Based Catalysis for PAA Activation

For water treatment, iron-based catalysts have been widely utilized owing to their environmental friendliness, high efficiency, and low cost. Among them, F e 2 + is one of the most extensively used iron catalysts, particularly because of its application in the Fenton process. In systems where PAA is activated by Fe(II)/Fe(III), the pH of the solution plays a critical role, similar to the traditional Fenton reaction. To prevent iron precipitation and maintain catalyst stability, acidic conditions must be carefully maintained [71]. In the first phase, the majority of Fe(II) is consumed to activate PAA and undergo oxidation to Fe(III). At a pH = 3, the performance of contaminant removal is heavily affected by the much lower reaction rate of Fe(III) with PAA (2.72 M 1 · s 1 ) [39], relative to that of Fe(II) with PAA (0.86–1.1 × 10 5   M 1 · s 1 ) [72]. Therefore, a major challenge for improving Fe(II)/PAA performance is to find efficient strategies to promote Fe(III) reduction back to Fe(II), considering the slow Fe(III)/Fe(II) redox cycle [73].
Several researchers have demonstrated the effectiveness of the Fe(II)/PAA system in removing CECs. For instance, Kim et al. [39] showed that the Fe(II)/PAA system could efficiently eliminate organic dyes, outperforming the conventional Fenton process. Yang et al. [63] reported that the Fe(II)/PAA system rapidly degraded p-arsanilic acid (5 µM), achieving 98% removal within just 20 s using 400 µM of PAA and 200 µM of Fe(II) at a pH of 3.0. In contrast, PAA alone achieved only 15% degradation after 45 min. Lin et al. [74] evaluated the degradation of diclofenac (DCF) by the Fe(II)/PAA system and observed that no significant elimination of DCF (10%) occurred under the conditions tested, which could be attributed to the low dosage of Fe(II) (1 μM) and the slow Fe(III)/Fe(II) redox cycling. To enhance the oxidation capacity of the Fe(II)/PAA process, they added 2,2′-azino-bis(3-ethylbenzothiazoline-6-sulfonate) (ABTS) as a reducing agent, achieving 92% of DCF removal within 30 min. ABTS effectively acted as an electron shuttle, accelerating the Fe(III)/Fe(II) redox cycle.
In recent years, iron-based heterogeneous catalysts have garnered considerable interest in PAA-driven AOPs. For example, Zhang et al. [66] used nano zero-valent iron for the activation of PAA, achieving approximately 95% elimination of tetracycline (TC) within 30 min. Moreover, Wang et al. [62] designed an innovative F e 2 + -modified zeolite/PAA system for the removal of sulfamethoxazole (SMX) at a neutral pH, achieving complete degradation within 60 min. Furthermore, Wang et al. [52] investigated the performance of the dual-metal catalyst, cobalt ferrite ( C o F e 2 O 4 ), combined with the PAA system for the removal of SMX as a target pollutant. The system achieved 87.3% elimination within 30 min under neutral conditions, using 200 μM of PAA and 0.1 g·L−1 of C o F e 2 O 4 . Furthermore, Wang et al. [75], applied pyrite tailings (PRT) to activate PAA for the removal of SMX. In situ Raman spectra confirmed the activation of PAA by PRT, with 93.67% of SMX (5 mg·L−1) removed within 50 min under near-neutral conditions (pH of 5.8). In this system, PAA activation was primarily attributed to leached Fe ions from PRT. Spectroscopic analyses and theoretical calculations revealed that sulfur sites in PRT enhanced Fe(II) regeneration, thus promoting the degradation process. Radical quenching experiments confirmed that acetylperoxyl and hydroxyl radicals were responsible for SMX degradation.

3.1.2. Cobalt-Based Catalysis for PAA Activation

Cobalt (Co) stands out as a highly efficient catalyst for radical generation, outperforming metals such as manganese, copper, and iron in PAA activation. This superior catalytic activity is attributed to the high standard redox potential of C o 3 + / C o 2 + ( E 0 = 1.82 V) and the ability of PAA to rapidly reduce C o 3 + to C o 2 + without external energy input [76].
The decomposition of PAA by Co leads to the formation of C H 3 C O O O and C H 3 C ( O ) O as the primary reactive radical species without the formation of H O . This process is illustrated by the reactions presented in Equations (14) and (15), which detail the generation of these reactive radicals [77].
C o 2 + + C H 3 C O O O H     C o 3 + + C H 3 C ( O ) O + O H
C o 3 + + C H 3 C O O O H     C o 2 + + C H 3 C O O O + H +
Kim et al. [20] found that the apparent second-order reaction rate constants for Co(II) and Co(III) with PAA (1.70 × 10 1 –6.67 × 10 2   M 1 · s 1 and 3.91 × 10 0 –4.57 × 10 2   M 1 · s 1 ), respectively, were comparable, based on kinetic simulations conducted over an initial pH range of 3.0–8.2. The comparable reaction rates of PAA with Co(II) and Co(III), along with the sustained Co(II)/Co(III) cycling, allow for the use of a low Co dosage in this AOP, improving the feasibility of this technology while minimizing concerns about Co toxicity. Several studies have examined the effectiveness of the C o 2 + /PAA process for various contaminants. For instance, Wang et al. [77] achieved approximately 90% degradation of SMX in 15 min through PAA activation using a minimal amount of Co (0.8 μM). The use of scavengers led to the conclusion that the organic radicals C H 3 C ( O ) O and C H 3 C O O O   played the primary role in degradation. Building on these findings, they proposed a reaction pathway outlining the involvement of RS in the Co/PAA system for eliminating the pollutant, as illustrated in Figure 2. Using the Co/PAA system, Kim et al. [20] conducted a comprehensive study on the elimination of 30 aromatic organic compounds, confirming C H 3 C O O O   as the primary active free radical.
On the other hand, a recent investigation by Liu et al. [41] revealed that under acidic conditions, the C o 2 + /PAA system can generate C o 4 + species, which may contribute to the degradation of organic contaminants. The pathway proposed for the formation of C o 4 + species is illustrated in Equations (16) and (17).
C o 2 + + C H 3 C O O O H     C o 2 + C H 3 C O O O H   (complex)
C o 2 + C H 3 C O O O H   ( c o m p l e x )     C o 4 + + C H 3 C ( O ) O + H +
However, unlike other metal/PAA systems, one of the advantages of C o 2 + /PAA is its ability to function under neutral conditions, which has led to the previously overlooked role of C o 4 + species. Neutral pH appears to favor the formation of oxygen-centered radicals, such as C H 3 C ( O ) O and C H 3 C O O O . Despite this, it warrants consideration, as certain pollutants, such as carbamazepine (CBZ), exhibit greater reactivity with C o 4 + than with oxygen-centered radicals [41].
Moreover, the Co(II)/PAA process faces challenges, such as limited reusability and risks of secondary pollution, which further restrict its applicability in PAA-based AOPs. In addition to these limitations, the potential ecotoxicological risks associated with Co leaching must also be carefully considered. The leaching of Co ions into the environment can have harmful effects on both aquatic and terrestrial ecosystems [78]. The toxicity of Co is highly influenced by its oxidation state and solubility, as well as various intrinsic and extrinsic factors [79]. Excessive exposure to Co is harmful to humans, potentially causing various diseases such as contact dermatitis, pneumonia, allergic asthma, and lung cancer [80], which limits its direct use in a homogeneous form.
Therefore, further research is needed to investigate the catalytic performance and mechanisms of cobalt-based heterogeneous systems. To overcome these drawbacks, cobalt-based catalysts can be engineered into heterogeneous forms, such as L a C o O 3 [69] and C o 3 O 4 [68], which have been designed to catalyze PAA. Among these, C o 3 O 4 particles exhibit outstanding catalytic performance, which is attributed to the coexistence of Co(II) and Co(III) within their structure. Nevertheless, certain pollutants, such as DCF, benzoic acid (BA), and CBZ, exhibit significant recalcitrance and greater resistance to the process [68]. In addition, zero-valent cobalt (ZVCo) has been developed as an excellent heterogeneous catalyst for PAA, targeting the removal of SMX as a contaminant. Zhou et al. [30] demonstrated that 99% SMX was removed by the ZVCo/PAA system within 5 min under neutral conditions. Additionally, ZVCo can be easily recycled due to its magnetic properties, which enhance its practicality and reusability in the system. Furthermore, even after four cycles, ZVCo maintained its remarkable ability to activate PAA for SMX degradation, highlighting its stability and reusability.
Due to its effectiveness in PAA activation, Co appears to be a promising option for the degradation of CECs. However, it is important to recognize that Co is associated with potential risks to human health. Therefore, Co should be neutralized before discharge in processes that utilize homogeneous Co, while heterogeneous systems must account for the potential leaching of Co into the sample. To mitigate these risks, several strategies can be implemented to minimize cobalt leaching in Co/PAA systems. These strategies aim to reduce Co release into the environment, enhance the efficiency of the catalytic process, and promote sustainable practices. Among these, cobalt-doped graphitic carbon nitride has shown promise in stabilizing ions, thereby preventing their release during catalytic reactions and improving the catalyst’s efficiency in pollution degradation [81]. Similarly, nano-island-encapsulated cobalt single-atom catalysts provide enhanced cobalt stability by embedding single cobalt atoms within a support matrix, further minimizing leaching while optimizing catalytic performance [82]. Additionally, chitosan-derived carbon has been explored to reduce cobalt leaching in disinfectant degradation, as its carbon matrix efficiently encapsulates cobalt-based nanoparticles, preventing metal release and maintaining catalytic efficiency [83]. These strategies contribute to improving the overall environmental impact of cobalt-based systems.

3.1.3. Manganese-Based Catalysis for PAA Activation

Manganese (Mn), a widely abundant transition metal, exhibits a broad range of oxidation states, from +2 to +7, in aqueous solutions and offers greater environmental friendliness compared to Co [84]. Unlike iron, it functions effectively across a wide pH range without generating sludge formation [85]. Mn(II) has been extensively studied as a homogeneous catalyst for activating various oxidants [86,87,88]. However, the role of soluble Mn(II) in mediating PAA activation remains insufficiently explored, likely due to its relatively low catalytic efficiency [89]. Consequently, there is limited information on the broader use of the Mn(II)/PAA system for eliminating organic pollutants. The only reported instance involves the successful degradation of the organic dye Orange II using Mn(II) salt in combination with PAA [43]. This study demonstrated that Orange II was degraded at a pH of 9.4; however, a high dosage of PAA ([PAA] = 5–20 mM, [Mn(II)] = 100 μM) was required. In contrast, degradation at acidic to neutral pH levels (pH 3–7) was significantly less effective under similar conditions. Further analysis revealed that Mn(IV), Mn(VI), and Mn(VII) were the primary reactive species generated during the reaction, as confirmed by UV–visible spectral changes and electron paramagnetic resonance (EPR) spectroscopy [43].
The limited performance of the Mn(II)/PAA system may stem from the inherent instability of high-valent Mn species, which are prone to spontaneous decomposition or disproportionation in water [90]. To overcome this challenge, the use of ligands in transition metal-based AOPs has been widely reported as an effective strategy for enhancing the degradation efficiency of organic pollutants [89,90,91]. By coordinating with transition metal ions, ligands form stable metal–ligand complexes that improve system stability and overall performance. Building on this principle, recent research has focused on the introduction of ligands. For example, Liu et al. [91] demonstrated the improved degradation of SMX by Mn(II)-activated PAA with the addition of nitrilotriacetic acid (NTA), which complexes with Mn(II) to form the Mn(II)-NTA complex, accelerating PAA activation. In the Mn(II)/PAA/NTA process, Mn(V) played a major role in SMX elimination. Notably, the Mn(II)/PAA/NTA system exhibited high treatment efficiency under acidic and neutral conditions, emphasizing its potential for wider environmental applications. Dong et al. [90] further found that the addition of ethylenediamine-N,N′-disuccinic acid (EDDS) significantly enhanced the elimination of atrazine (ATZ) in the Mn(II)/PAA process. Moreover, in the Mn(II)/EDDS/PAA system, Mn(V) was confirmed as the primary reactive species responsible for ATZ removal, as the inclusion of the ligand improved the stability of manganese intermediates. Consequently, the role of high-valent metals (HVMs)—transition metals in oxidation states higher than +3, capable of facilitating strong oxidative transformations—in transition metal-catalyzed PAA systems is of considerable importance and cannot be overlooked. For instance, Kim et al. [89] reported that Mn(II) alone exhibits limited reactivity with PAA; however, the addition of picolinic acid (PICA) significantly enhances the decomposition of PAA by Mn(II). The PAA-Mn(II)-PICA system efficiently removes various contaminants, including bisphenol A (BPA), NAP, SMX, CBZ, and trimethoprim, at neutral pH, achieving over 60% removal within 10 min in both pure water and wastewater environments.
However, the incorporation of organic ligands to enhance the Mn(II)/PAA process may elevate chemical oxygen demand levels or lead to the formation of by-products through reactions with disinfectants. Moreover, their use adds chemical complexity to the wastewater, requiring additional removal steps. To improve the oxidation potential of the Mn(II)/PAA process, it is crucial to investigate more economical and eco-friendly approaches.
To improve the oxidation potential of the Mn(II)/PAA process, it is crucial to investigate more economical and eco-friendly approaches. Several strategies can be evaluated, including the use of green ligands to mitigate the environmental burden on wastewater and the exploration of the feasibility of recovering organic ligands post-use.

3.2. UV Irradiation-Induced Activation of PAA

The activation of PAA by UV irradiation is among the most extensively studied processes, as UV radiation initially induces the homolytic cleavage of the O-O bond in PAA, generating H O and C H 3 C ( O ) O (Equation (18)) [40,92,93]. For this purpose, the most commonly used UV light wavelength is 254 nm, which is emitted by low-pressure mercury lamps. Zhang et al. [40] quantified the quantum yield for PAA decomposition at a UV of 254 (Φ = 0.88 ± 0.04 mol-Ein−1), which exceeds that of H 2 O 2 (Φ = 0.5 mol−1), indicating a faster rate with a comparable amount of absorbed light. Since H 2 O 2 always presents in equilibrium with PAA solutions, it also participates in the degradation processes under ultraviolet radiation, as shown in Equation (19).
C H 3 C O O O H   + h v     C H 3 C ( O ) O   + H O
H 2 O 2   +   h v     2 H O
For the degradation of CECs, numerous recent scientific publications have focused on utilizing the UV/PAA process, and these findings are summarized in Table 2.
Notably, UV light alone can directly degrade certain organic contaminants that have dissociative absorption bands in the UV range [94]. However, in most cases, achieving the direct UV mineralization of organic pollutants is challenging. Nevertheless, when PAA is combined with UV irradiation, the CECs can be destroyed by various mechanisms, offering a potential solution to this challenge: first, direct oxidative degradation, utilizing the strong oxidizing property of PAA; second, direct photolysis, where organic pollutants absorb photons from ultraviolet light and undergo degradation; and third, the well-known activation of PAA through UV irradiation, leading to the production of highly active radicals such as H O and R O . The presence of R O in the UV/PAA system is thought to be responsible for its greater selectivity compared to the UV/ H 2 O 2 process [53]. A number of studies have shown that the UV/PAA system can improve the efficiency of pollutant degradation through a synergistic effect, suggesting that UV/PAA is a promising technology for water treatment. For instance, Sharma et al. [95] reported degradation rates ranging from 80% to 100% for various chlorophenols. Their findings indicate that the position and number of chlorine atoms in the aromatic ring structure play a significant role in the process. Similarly, Cai et al. [21] investigated the effectiveness of the UV/PAA process in degrading several pharmaceutical compounds. They demonstrated that, at a pH of 7.1, while PAA alone resulted in minimal degradation, all compounds were degraded by more than 93.5% with 1 mg·L−1 of PAA under UV irradiation, further highlighting the effectiveness of the UV/PAA process.
To identify the reactive species responsible for the degradation, they also conducted scavenging experiments. They stated that the removal was primarily driven by H O oxidation, succeeded by oxidation by other radicals, mainly direct photolysis and organic radicals [40]. Additionally, organic radicals reacted rapidly with NAP but not with ibuprofen or carbamazepine, demonstrating their greater selectivity and suggesting a higher affinity for specific pollutants [40].
Multiple factors, such as PAA dosage and pH, have been identified as influencing the performance of the UV/PAA system. Among these factors, UV exposure at 254 nm plays a crucial synergistic role, as it is a light-activated system, as demonstrated by multiple studies [21,40]. On the other hand, radical generation in solution is constrained by the system’s photon absorption capacity [93]. Nonetheless, based on the available literature, the impact of the reactor design and configuration on the UV/PAA system remains largely unexplored.
Table 2. A summary of studies using UV irradiation for PAA activation in the removal of CECs.
Table 2. A summary of studies using UV irradiation for PAA activation in the removal of CECs.
CEC[CEC]0Removal Efficiency (%)ActivatorPAA DosepHReaction TimeRef.
Fluoxetine5 mg·L−1100UV (254 nm) irradiation: (647–3502) W·m−35–100 mg·L−1730 min[53]
Sulfamethoxazole5 mg·L−1100UV (254 nm) irradiation: (647–3502) W·m−35–100 mg·L−1730 min[53]
Diclofenac1 μM90UV (254 nm) intensity: 2.12 × 10−6 E·L−1·s−11 mg·L−17.1<5 min[21]
Ibuprofen1 μM90UV (254 nm) intensity: 2.12 × 10−6 E·L−1·s−11 mg·L−17.130 min[21]
Carbamazepine1 μM>90UV (254 nm) intensity: 2.12 × 10−6 E·L−1·s−11 mg·L−17.130 min[21]
Naproxen1 μM>95UV (254 nm) intensity: 2.12 × 10−6 E·L−1·s−11 mg·L−17.110 min[21]
Chloramphenicol25 mg·L−1100UV (254 nm) doses: 0~12.5 W·m−2 5–50 mg·L−1Not mentioned120 min[96]
Clofibric acid1 μM>90UV (254 nm) intensity: 2.12 × 10−6 E·L−1·s−11 mg·L−17.110 min[21]
Venlafaxine5 mg·L−1100UV (254 nm) irradiation: (647–3502) W·m−35–100 mg·L−1730 min[53]
4-chlorophenol4 μM>90UV (254 nm)3042 mg·L−19.55 min[54]
Pentachlorophenol4 μM100UV (254 nm)3042 mg·L−19.55 min[54]
2,4,6-trichlorophenol4 μM>80UV (254 nm)3042 mg·L−19.55 min[54]
2,4-dichlorophenol4 μM100UV (254 nm)3042 mg·L−19.55 min[54]
Bezafibrate1 μM80UV (254 nm) intensity: 2.12 × 10−6 E·L−1·s−11 mg·L−17.1120 min[21]
Carbamazepine5 mg·L−1100UV (254 nm) irradiation: (647–3502) W·m−35–100 mg·L−1730 min[53]
Naproxen4 μM80UV (254 nm) intensity: 9.04 × 10−8 E·L−1·s−120 mg·L−1714 min[92]

3.3. Carbon-Based Catalyst for the Activation of PAA

Green, metal-free catalysts mark significant advancements in modern chemistry, with their widespread adoption offering great promise due to their ability to completely eliminate toxic metal leaching and secondary water contamination [59]. Among these, carbon materials stand out with remarkable advantages in terms of stability, activity, and renewability in comparison to traditional metal or metal oxide catalysts. Additionally, their cost-effectiveness and environmental friendliness position them as compelling alternatives to conventional catalysts, aligning with the goals of sustainable development [60].
Carbon materials occur in various allotropic forms, including activated carbon [59], carbon nanotubes [97], biochar [60], and graphene [98]. In the field of water treatment, carbon materials not only serve as adsorbents with excellent capacity but also function as catalysts to enhance PAA activation for the removal of organic pollutants. Table 3 presents selected studies employing carbon materials for PAA activation.
This capability was first demonstrated in 2015 by Zhou et al. [99], who combined activated carbon fibers (ACFs) with PAA for the removal of the dye Reactive Brilliant Red X-3B, achieving nearly 97% removal after 45 min. Similarly, in 2022, another example concerned the use of carbon nanotubes (CNTs), where pristine CNTs were utilized to activate PAA and characterize the reactive species and associated active sites within the CNTs [97].
The dominant species responsible for the elimination of BPA was the CNT-PAA* complex, while a secondary role was played by adsorbed H O   radicals. Additionally, the adsorption energy of PAA on the CNT edge was stronger than that on the CNT matrix, as shown by density functional theory calculations, suggesting a favorable interaction between CNT edges and PAA, which facilitates PAA activation for adsorbed H O generation.
Furthermore, biochar is noted for its durability, insolubility, and high carbon content [100]. Its surface contains numerous defect structures and oxygen-containing functional groups, which are capable of donating electrons to oxidants and producing free radicals [101]. In a study by Hou et al. [102], they investigated the activation of PAA using biochar derived from chicken manure for the degradation of SMZ. Their study demonstrated that R O was identified as the dominant reactive substance, which was produced via electron transfer. In another study, Zhang et al. [103] explored the oxidation of SMZ using PAA activated by biochar, which was thermally modified under neutral environmental conditions. Their EPR and free radical quenching experiments clearly demonstrated the existence of 1 O 2 in the system, which plays a dominant role in the degradation of SMZ. One such material, graphene, is composed of a two-dimensional framework, formed by the organization of carbon atoms in a honeycomb pattern via sp2 bonding [104]. Due to its unique structure, graphene exhibits hydrophobic characteristics, which pose challenges for its application in AOPs. However, researchers have made considerable advancements by investigating graphene-based derivatives, such as graphene oxide, reduced graphene oxide, and other more complex synthetic materials. These derivatives exhibit their adaptability in AOPs [105], particularly when oxygen-containing functional groups, such as hydroxyl, carboxyl, and carbonyl groups, are present. These groups reduce the hydrophobicity of GO, significantly enhancing its reactivity in AOPs [106]. Additionally, GO can promote mediated electron transfer through its intercalated structure, further improving catalytic removal efficiency in AOPs [107]. In this context, Tshangana et al. [108] utilized the graphene oxide quantum dots/PAA system for the removal of sulfasalazine, demonstrating its effective degradation. To reduce the insulating property of GO, rGO is engineered from it [109]. As a result, rGO displays a greater density of defects and a notably increased specific surface area in comparison to the original material [110]. Furthermore, rGO facilitates the production of stable intermediates by activating oxidizing agents due to its distinctive single-layer structural arrangement [105].
Table 3. A summary of studies employing carbon materials for PAA activation in CECs removal.
Table 3. A summary of studies employing carbon materials for PAA activation in CECs removal.
CECCatalystProcesses[CEC]0Removal Efficiency (%)PAA DosePHReaction TimeRef.
Reactive Brilliant Red X-3BACFsACFs/PAA50 μM975 mM745 min[99]
sulfamethoxazolerGOrGO/PAA10 μM1000.1 mM55 min[98]
bisphenol ACNTsCNT/PAA0.02 mM96.40.25 mM720 min[97]
phenolCNT950CNT950/PAA20 μM1000.15 mM1010 min[111]
sulfamethoxazoleAC600AC600/PAA20 mg·L−199.40.26 mM7150 min[59]
sulfamethazineactivated biochar (ABC)ABC/PAA5 mg·L−172.80.07 mM7100 min[103]
phenolcarbonized polyaniline (CPANI)CPANI/PAA10 μM970.1 mM760 min[112]
The degradation mechanisms of the carbon material/PAA systems are currently classified into three main types:
-
The first mechanism involves the generation of singlet oxygen 1 O 2 , generated through PAA activation by N-doped carbonaceous catalysts. For example, Tian et al. [112] employed carbonized polyaniline (CPANI) as the catalyst, where the C=O group on CPANI activated PAA to generate 1 O 2 , the main reactive species driving phenol degradation.
-
The second mechanism is based on direct electron transfer (DET). In this context, Kong et al. [98] demonstrated the effectiveness of reduced graphene oxide (rGO) in activating PAA for the rapid removal of SMX, achieving a near-complete elimination of the pollutant within just 2 min. Through a combination of quenching experiments, open-circuit potential measurements, and probe-based studies, they confirmed that DET was the dominant degradation pathway. The rGO/PAA system exhibited strong removal performance even in complex water matrices, highlighting the advantages of DET-driven oxidation. Building upon these findings, subsequent research explored other carbon-based materials to enhance DET processes. For instance, Kong et al. [111] emphasized the role of the physicochemical properties of CNTs on organic pollutant removal and PAA activation. The CNT/PAA system was also dominated by the DET oxidation pathway, achieving high MP removal rates. The enhanced catalytic efficiency of surface-regulated CNTs was attributed to reinforced DET, driven by the increased oxidative potential of the CNT/PAA complex and the enhanced electrical conductivity of CNTs. Furthermore, the larger specific surface area and lower oxygen content of CNTs were found to contribute significantly to the elevated oxidative potential, with electrical conductivity closely linked to their degree of graphitization.
-
The third degradation mechanism involves the participation of active radicals, as demonstrated by a study on the degradation of 4-chlorophenol (4-CP) via PAA activation using a newly synthesized biochar (P5S5-SDBC) [60]. Mediated electrochemical oxidation and reduction tests revealed that persistent free radicals (PFRs), generated by structural defects in the biochar, acted as electron shuttles, enhancing PAA activation and promoting the formation of reactive species. Electron paramagnetic resonance (EPR) measurements further confirmed that R O radicals were the primary active species responsible for 4-CP degradation [60].
In recent years, PAA-AOPs, particularly those utilizing carbon-derived catalysts, have attracted considerable attention in environmental remediation. However, carbon materials often suffer from limited stability, primarily due to their vulnerable edges, surface defects, and susceptibility to interactions with reaction intermediates through π–π stacking or hydrogen bonding [113]. To address these challenges, the strategy of encapsulating metal cores within robust carbon shells has emerged as a promising solution. This approach effectively protects the internal metal centers, enhances catalyst stability, minimizes metal leaching, and reduces the risk of secondary pollution. Despite these advances, research into the use of metal–carbon composite materials for contaminant removal via PAA activation remains relatively limited and warrants further exploration.

3.4. Comparative Analysis of Catalytic System for PAA Activation: Transition Metals, UV Irradiation, and Carbon-Based Catalysts

Various catalytic systems, such as transition metals, UV irradiation, and carbon-based catalysts, have been explored to improve the activation of PAA. Transition metal catalysts are widely utilized due to their ability to generate ROS, including C H 3 C O O O and C H 3 C ( O ) O . They are cost-effective and abundant, making them an attractive option for various applications. Furthermore, systems using transition metals are straightforward to implement, simplifying operational procedures. However, their long-term stability and activity may decrease over time due to factors like oxidation, which limits the overall effectiveness and longevity of the catalyst. The use of transition metals in catalytic processes can lead to secondary pollution due to the leaching of metal residues. This may result in the accumulation of harmful metal ions in treated water, posing environmental risks. In homogeneous catalytic systems, it is challenging to recover transition metal catalysts once they have been used in the reaction. This complicates the recycling process and increases the overall operational costs. In contrast, the UV irradiation activation of PAA does not require any chemical additives, making it an environmentally friendly option that minimizes the risk of by-product generation. The simplicity of operating UV systems adds to their appeal. However, UV activation is energy-intensive, which can increase operational costs. Additionally, UV may struggle to penetrate turbid or colored water, thereby reducing its effectiveness. Carbon-based catalysts are known for their environmental compatibility, as they are non-toxic and eco-friendly. They offer a high surface area for both catalytic reactions and adsorption, which enhances their performance. Another advantage is their reusability, contributing to sustainability in long-term applications. On the downside, carbon-based catalysts are typically less efficient in activating PAA compared to metal-based catalysts. The cost of producing high-quality carbon materials can be prohibitive, and some carbon materials require modifications to increase their reactivity for optimal performance.
Each catalytic system presents distinct advantages and limitations. Selecting the appropriate process depends on specific treatment objectives, environmental considerations, and economic factors.
The proficiency of PAA-based AOPs in the remediation of CECs is closely linked to the nature of the RS involved. PAA, upon activation, generates highly RS, such as H O , C H 3 C ( O ) O O , 1 O 2 , and HVMs, all of which play pivotal roles in the oxidative elimination of CECs. The determination of these RS is essential for understanding the underlying mechanisms of PAA-based AOPs, as their generation directly influences the efficiency and selectivity of pollutant degradation. Through techniques such as scavenging experiments, isotope labeling, and spectroscopy, researchers can identify and quantify these RS, providing crucial insights into optimizing operational conditions and improving the overall performance of PAA-based AOPs in environmental remediation.

4. Determination of Reactive Species in PAA-Based AOPs for Water Decontamination

The efficiency of CEC treatment predominantly depends on the RS generated during the PAA activation process. These include free radicals such as H O [46], C H 3 C O O ,   C H 3 C ( O ) O O , C H 3 O O ,   C H 3 C O , and C H 3 [50,66], as well as non-radical species like 1 O 2 [103], HVMs like Fe(IV) and Co(IV) [114] (Figure 3), electron-transfer processes [98], and direct oxidation [36].
To provide deeper insights into the degradation mechanisms of organic pollutants in PAA-based AOPs, it is crucial to identify the primary RS driving the degradation process. Based on prior research, C H 3 has been identified as a less significant radical in pollutant degradation due to its limited oxidizing ability and its propensity to react with dissolved oxygen, forming C H 3 O O [19]. Notably, the oxidizing strength of C H 3 O O is comparable to that of C H 3 , further diminishing its effectiveness in degrading pollutants.
Additionally, the spontaneous decomposition rate constants of C H 3 C ( O ) O , as determined via Equations (5) and (6), are 1.0 × 10 9   M 1 · s 1 and 2.3 × 10 5   M 1 · s 1 , respectively. It is worth noting that its spontaneous decomposition rate typically exceeds its reaction rate with pollutants [66], which may limit its direct contribution to degradation. However, this behavior aligns with previous studies indicating that H O , along with other oxygen-centered radicals, such as C H 3 C ( O ) O O , play crucial roles in oxidative degradation pathways [20,21,35,40,115].

4.1. Free Radicals

4.1.1. Organic Radicals: C H 3 C ( O ) O and C H 3 C ( O ) O O

Prior studies have demonstrated that the C=C double bond is highly prone to attack by organic radicals. For instance, the reaction rate constant between R O and the C=C double bond in β-carotene ( C 40 H 56 ) is 9.2 × 10 8   M 1 · s 1 , indicating that β-carotene can effectively neutralize R O [116]. Similarly, 2,4-hexadiene (2,4-HD), which also contains a C=C double bond, is widely recognized for its ability to scavenge R O , with a rate constant of approximately 1 × 10 10   M 1 · s 1 for the reaction. Furthermore, due to its fast reaction with H O (k2,4-HD,•OH = 9.2 × 109  M 1 · s 1 ), (2,4-HD) can simultaneously scavenge both R O and H O . Consequently, the inhibitory effect of tert-butyl alcohol (TBA), which is typically a scavenger of H O and 2,4-HD, on compound degradation may vary, and this variation can demonstrate the role of R O in the system [59,66]. In this context, Zhang et al. [117] revealed that M n 2 + can act as a specific scavenger for C H 3 C ( O ) O O in PAA-based AOPs via electron transfer, without significantly influencing the decomposition of PAA itself [117].
The core concept behind using the quenching method to identify a specific active species is that the quencher must not interfere with the interactions between pollutants and other compounds within the system. For instance, in the CNT/PAA system, introducing 2,4-HD markedly reduces the adsorption of BPA by CNTs [97]. Even at minimal concentrations, 2,4-HD induces the aggregation of CNT powder in the solution, thereby affecting the overall reaction conditions of the system. To mitigate these effects, bromotrichloromethane emerges as a novel alternative, effectively terminating R O without altering the catalyst’s adsorption properties, making it the most suitable scavenger available [97,118].
On the other hand, for the identification of organic radicals, other methods such as the probe technique can be used. This approach relies on introducing trace amounts of probe compounds, allowing chemical probe experiments to determine the type and presence of active species based on the degradation of the probe molecules and the characteristics of the resulting products. For example, 2,2,6,6-tetramethyl-1-piperidinyloxy (TEMPO) is a stable nitroxide radical that acts as a radical probe for detecting C H 3 and C H 3 C ( O ) O . The generated methyl-TEMPO ( C H 3 -TEMPO) can be analyzed using mass spectrometry [52]. For instance, during the pyrite/PAA process, the generation of C H 3 C ( O ) O -TEMPO is evidenced by the C H 3 C ( O ) O -TEMPO signal (m/z 216.1598 and a retention time of 1.239 min) [71].
Additionally, the EPR method is frequently utilized to accurately detect the presence of R O . 5-diisopropoxyphosphoryl-5-methyl-1-pyrroline N-oxide (DIPPMPO) serves as a widely accepted spin-trapping agent. The extracted hyperfine splitting constants (aP, aN, and aH) of the captured radicals are then simulated using Easyspin software (version 5.2.28 toolbox for MATLAB (The MathWorks, MA, USA)), which also aids in analyzing the resulting fitting curve. This process confirms the presence of C H 3 C ( O ) O , C H 3 , and C H 3 O O [119,120]. However, the detection of C H 3 C ( O ) O O remains inconclusive, likely due to its high instability and the absence of relevant fitting information [98]. Recently, Chen et al. [121] conducted a comprehensive study on radical identification through in situ EPR analysis, successfully establishing a reliable method for detecting C H 3 C ( O ) O O in PAA-based AOPs by trapping it with DMPO in ethanol under a nitrogen atmosphere.

4.1.2. Inorganic Radicals: H O

In the context of PAA-based AOPs, H O   represents the principal inorganic reactive species, noted for its exceptionally high reactivity, with reaction rate constants for different compounds typically in the range of  10 8 10 10   M 1 · s 1 [122]. Within such oxidative environments, alcohols are widely acknowledged as potent radical scavengers. For instance, methanol (MeOH), ethanol (EtOH), and tert-butyl alcohol (TBA) effectively scavenge H O , preventing radical-induced reactions, with respective rate constants of 9.7 × 10 8   M 1 · s 1 , (1.2–2.8) ×   10 8   M 1 · s 1 , and (3.8–7.6) ×   10 8   M 1 · s 1 [122,123]. Nevertheless, MeOH and EtOH are prone to reacting with organic radicals, subsequently generating aldehydes or alcohol-derived radicals [55]. Conversely, in PAA-based AOP systems, TBA uniquely acts as a highly selective scavenger for   H O , as clearly depicted in Figure 4. The substantial decline in pollutant degradation following the introduction of TBA strongly underscores the indispensable contribution of H O within these oxidative systems [21,55].
In addition to alcohols, several other compounds can be effectively utilized to detect the presence of H O radicals in AOPs. For example, nitrobenzene (NB) and p-chlorobenzoic acid (p-CBA) rapidly react with H O , exhibiting reaction rate constants of 3.9 ×   10 9   M 1 · s 1 and 5.0 ×   10 9   M 1 · s 1 , respectively [55,77]. Additionally, these compounds demonstrate negligible reactivity toward other reactive species, rendering them highly selective and dependable probes for H O detection. Consequently, the measurable reduction in the concentrations of these probe compounds serves as strong evidence confirming the presence and activity of H O .
On the other hand, in certain heterogeneous systems, commonly employed alcohols may not completely quench surface-bond H O due to substantial electrostatic repulsion exerted by the catalyst surface. Given its high reactivity, H O reacts with potassium iodide (KI) at a rate constant of 1.2 ×   10 10   M 1 · s 1 [97,124], and KI has been considered as an alternative probe for detecting surface-bound H O in such a system. Moreover, sodium azide ( N a N 3 ) is also recognized as an efficient quencher of surface-bound radicals, with a reaction rate constant of 1.2 ×   10 10   M 1 · s 1 for H O [125]. Notably, its low dielectric constant (4.5), compared to EtOH (25.3) and TBA (12.5), enhances its suitability for quenching surface-bound OH in heterogeneous systems [126]. It is relevant to point out that KI and N a N 3 act as powerful electron donors, capable of directly reacting with PAA, leading to its dissipation and subsequently hindering the elimination of model compounds [127]. Therefore, when using KI or N a N 3 as quenchers, the careful monitoring of changes in PAA concentration is essential. To offset potential depletion, the supplemental addition of PAA into the system is recommended during quenching experiments.

4.2. Non-Radical Species

4.2.1. Singlet Oxygen (1 O 2 )

1 O 2 is a highly selective oxidant compared to other inorganic radicals. It can react with unsaturated organic compounds through electrophilic attack or by abstracting electrons [122]. 1 O 2 exhibits a lower redox potential (0.81 V) compared to H O , indicating that its reaction rate with most organic compounds is generally slower than that of radicals [128]. The reaction rate of 1 O 2 with organic compounds typically ranges from 10 4 to 10 7   M 1 · s 1 . According to earlier research, sites that favor the production of 1 O 2 comprise graphitic N, electron-deficient carbon, N-metal sites, oxygen vacancies, and ketone groups [129,130]. In PAA-based AOPs, the dissociated PAA anion C H 3 C O O O can react with the central carbon atom of PAA to produce 1 O 2 , as shown in Equation (20) [131].
C H 3 C O O O H + C H 3 C O O O     C H 3 C O O H   + C H 3 C ( O ) O + 1 O 2
A widely used scavenger for 1 O 2 is furfuryl alcohol (FFA), which is acknowledged as a highly effective quencher. The rate constant for the reaction between FFA and 1 O 2 in aqueous solutions is k F F A ,   1 O 2 = 8.3 × 10 7   M 1 · s 1 [59,132]. The steady-state concentration of 1 O 2 is calculated using the pseudo-first-order degradation rate constant ( K o b s ) with FFA as the target contaminant (Equation (21)) [133].
K o b s   F F A = d   F F A d t = k F F A ,   1 O 2 × F F A × 1 O 2   1 O 2 S S
On the other hand, FFA reacts effectively with H O , with a reaction rate constant of 1.5   × 10 10   M 1 · s 1 . Therefore, excess TBA must be added during the quenching process if H O is generated [59]. Wang et al. [55] reported that CBZ reacts rapidly with 1 O 2 as well as H O . To scavenge H O , excess TBA is used, and the removal of CBZ serves as a clear indicator of the existence of 1 O 2 . Additionally, the reaction rate of β-carotene with 1 O 2   ( 2 3 × 10 10   M 1 · s 1 ) highlights its role as an effective 1 O 2 scavenger [134].
In EPR measurements, 2,2,6,6-tetramethyl-4-piperidinol (TEMP) is commonly selected as a trapping agent for 1 O 2 . Its oxidation by 1 O 2 leads to the formation of TEMPO, which produces a distinct triplet signal. For instance, TEMPO can be detected in the thermoactivated PAA process, with its production rate increasing as the temperature elevates [55].
Moreover, prior studies have demonstrated that the lifetime of 1 O 2 in heavy water (deuterium oxide, D 2 O ) ranges from 20 to 32 μs, which is more than tenfold longer than in H 2 O (approximately 2 μs), thereby extending the reaction window. This prolonged lifetime enhances the degradation efficiency of pollutants mediated by 1 O 2 . The increased pollutant removal observed in D 2 O further underscores the significant role of 1 O 2 in oxidative processes [135].
Furthermore, numerous studies have reported that 1 O 2 serves as a major reactive species for degrading organic compounds in H 2 O solvents [136,137]. However, these conclusions require stronger validation through additional experiments. Notably, water itself can act as a scavenger of 1 O 2 , with a reaction rate constant of 10 5   M 1 · s 1 . Since water typically serves as the medium and exists in concentrations much higher than those of organic pollutants in aqueous environments, it is hypothesized that 1 O 2 contributes significantly to the degradation of organic compounds. For this to occur, the reaction rate between 1 O 2 and pollutants must exceed 10 8   M 1 · s 1 , based on competitive kinetics.
Given these considerations, D 2 O is recommended for use in PAA-based AOPs to leverage its ability to extend the lifetime and effectiveness of 1 O 2 .

4.2.2. High-Valent Metals (HVMs)

HVMs generated during PAA activation have recently been recognized as key intermediates [130]. Unlike radicals, HVMs exhibit longer lifetimes, potentially making them more stable and effective under certain conditions [138]. In contrast, radical-mediated oxidation processes are often impeded by the presence of natural organic matter (NOM), including oxyanions and halide ions, which are commonly found in water matrices and significantly reduce their efficiency [139,140]. Similarly, 1 O 2 exhibits substrate-specific reactivity [141] and relatively low oxidation capability [128]. This has led to significant debate regarding whether 1 O 2 serves as the primary reactive oxygen species responsible for pollutant degradation during AOPs [122]. Conversely, HVMs exhibit strong oxidation potential, as indicated by E 0 ( F e I V O 2 + / F e 3 + ) = 2   V [142], C o I V O 2 + = 1.81   V [143], and C u I I I O 2 + = 1.7   V [144]. Their effectiveness in contaminant removal is influenced by the presence of anions, NOM, and alcohols [41,145]. As a result, HVMs have attracted growing research interest and ongoing scientific debate over the past decade [138,146]. A key advantage of HVMs lies in the efficient generation of high-valence Fe species, which are produced from their corresponding low-valence forms (Fe(II) and Fe(III)) via a single-step, two-electron transfer process via Equation (22) [143]. Notably, compared to H 2 O 2 , PAA more effectively promotes the oxidation of low-valent iron to Fe(IV), due to its lower energy of the lowest unoccupied molecular orbital (LUMO) and weaker O–O bond strength, both of which facilitate enhanced electron transfer [39].
C H 3   C O O O H +   F e ( I I )     C H 3 C O O H + F e I V O 2 +
HVMs produced during PAA activation have been widely identified using a combination of probe compounds, such as dimethyl sulfoxide (DMSO) and methyl phenyl sulfoxide PMSO, with 18O-labeled H 2 O ( H 2 18O) [143,147], alongside various spectroscopic techniques. These identification methods have provided evidence that HVMs form in both homogeneous and heterogeneous PAA activation processes.
In 2019, Kim et al. [39] were the first to propose that Fe(IV) may play a pivotal role in the degradation of certain CECs, including NAP and BPA, within the Fe(II)/PAA system. This hypothesis was based on the substantial conversion of PMSO to PMSO2 and the minimal quenching effect observed in the presence of TBA at pH levels of 3, 5, and 7. Despite advances in the understanding of HVMs, the potential use of Co(IV) for wastewater remediation remained largely unexplored and unclear until 2021. In that year, Liu et al. [41] confirmed the formation of Co(IV) during PAA activation by Co(II) through PMSO-based probe tests, 18O isotope labeling experiments, and in situ Raman spectroscopy. This finding aligns with several studies that have highlighted the pivotal function of HVMs. For instance, in the F e 2 + /PAA/UVA process, the high-valent iron (IV)–oxo complex ( F e I V O 2 + ) was identified as the dominant RS. This species plays a critical role in the efficient and selective degradation of CECs. The research demonstrated that F e I V O 2 + serves as the primary oxidant, driving the degradation process, while the generation of organic radicals and H O was found to be minimal in comparison [148]. Similarly, in the cobalt(II)/PMS system, the formation of C o 4 + plays a pivotal role in AOPs, further reinforcing the importance of HVMs as primary oxidants. It was observed that PMSO was readily oxidized to the corresponding sulfone (PMSO2) with a transformation ratio of ∼100% under acidic conditions, which strongly implied the generation of high-valent cobalt–oxo species C o 4 + instead of S O 4   in the Co(II)/PMS process [149].
The interaction between HVMs and pollutants mainly occurs through three fundamental processes: hydrogen atom extraction [150], which involves the removal of a hydrogen atom from the pollutant molecule, frequently generating reactive intermediates; electron transfer [146], where the metal acts as an oxidizing agent by acquiring electrons from the pollutant; and oxygen atom transfer [151], whereby the metal aids in incorporating an oxygen atom into the pollutant, leading to its oxidation.
For the hydrogen atom extraction process involving aliphatic alcohol, Fe(IV) can remove the hydrogen atoms from the hydroxyl group, leading to the formation of an aldehyde or ketone, while Fe(II) is simultaneously formed through a two-electron transfer from Fe(IV). Additionally, the H atom on the α-C of RnCH–OH can be abstracted to form the radical intermediate R2C•–OH, which subsequently reacts with O 2 . The formed RC(OH)OO• intermediate undergoes an elimination reaction to produce R2C=O [150]. In the context of electron transfer, HVMs can react with compounds containing electron-donating groups via electrophilic addition pathways. For instance, Fe(IV) reacts with phenol by initially attacking the para-carbon of the aromatic ring. This electrophilic addition, followed by subsequent elimination, leads to the formation of hydroquinone [146].
Moreover, through an oxygen atom transfer mechanism, HVMs can oxidize sulfoxides into their corresponding sulfones. This pathway is particularly distinct, as sulfones are generally not produced in radical-driven reactions involving sulfoxides, highlighting the unique reactivity of HVMs in such transformations [152].

5. By-Products and Hazard Assessment of PAA-Based AOPs

Through the formation of RS, PAA-based AOPs have attracted increasing interest due to their efficiency in breaking down CECs. However, during oxidative reactions in water treatment, several by-products can be produced, some of which require assessment to ensure environmental safety. Among these, the primary by-products include carboxylic acids such as acetic acid [46]. The accumulation of acids in treated water can significantly lower pH, thereby affecting water quality and increasing the need for post-treatment neutralization.
Long-term acid accumulation can adversely affect aquatic ecosystems, impairing respiration and ion regulation in organisms, reducing biodiversity, and favoring acid-tolerant species. Chronic exposure may also result in sublethal effects, such as impaired reproduction in invertebrates, behavioral changes in fish, and reduced photosynthetic efficiency in algae. At the microbial level, while low concentrations of acetic acid may stimulate heterotrophic activity, elevated levels can inhibit sensitive populations and disrupt key processes such as nitrification. Moreover, due to its high biodegradability, acetic acid may increase biochemical oxygen demand (BOD), contributing to oxygen depletion in stagnant or eutrophic environments. When reused for irrigation, residual acetic acid may alter soil pH and hinder plant growth, particularly in alkaline or sensitive soils. Although acetic acid is not considered highly hazardous, it is regulated in effluent discharge standards to mitigate ecological acidification and oxygen stress.
In addition to acids, PAA disinfection also results in the formation of oxygenated by-products such as aldehydes and ketones, whose formation is directly related to the PAA dosage [153]. These compounds are biologically reactive and may persist under suboptimal degradation conditions. Formaldehyde, for example, is a probable human carcinogen (IARC Group 2A) and presents significant ecotoxicological risk [154]. It has an acute 96 h LC50 of 24 mg/L for Pimephales promelas and Oncorhynchus mykiss, and a LOAEL of 15 mg/kg/day in rodent studies, indicating both irritant and developmental effects [155]. Acetaldehyde, while of moderate toxicity, has an LC50 of 122 mg/L for Daphnia magna and has been associated with developmental toxicity under chronic exposure. Both aldehydes are capable of inducing oxidative stress and potential endocrine disruption and may bioaccumulate to a limited extent.
Importantly, aldehydes like formaldehyde may undergo further oxidation in the presence of excess PAA, forming carboxylic acids such as formic acid, introducing secondary environmental impacts [153].
Furthermore, in natural waters containing halide ions (e.g., Br and Cl), these can be oxidized by PAA into halogenated species such as hypohalous acids and halogen radicals [156]. These reactive intermediates can then react with organic matter to form halogenated by-products. Bromide ions are particularly concerning due to the significantly higher reaction rate constant with PAA compared to chloride [157]. Brominated organic by-products, such as bromoform, are persistent, bioaccumulative, and highly toxic. For instance, bromoform has an LC50 of approximately 5 mg/L for Daphnia magna and has been linked to reproductive and developmental toxicity in fish embryos.
To minimize the risks of secondary contamination, a comprehensive understanding of the formation mechanisms and influencing factors—such as PAA concentration, water matrix composition (especially halide content), activation method, and pH—is essential. Optimizing operational parameters and monitoring by-product profiles are crucial for ensuring the environmental safety of PAA-based AOPs over both short and long timescales.

6. Factors Affecting the Removal of CECs in PAA-Based Systems

6.1. Impact of Initial PAA Concentration

The initial dosage of PAA is a key factor affecting the removal efficiency of CECs in PAA-based processes. Generally, increasing the PAA concentration enhances the generation of RS in solution, thereby accelerating the degradation rate of CECs. Furthermore, higher PAA dosages increase the availability of oxidants. However, this benefit typically reaches a maximum at an optimal concentration, after which excessive PAA may actually hinder degradation performance. An excessive amount of PAA may act as a scavenger of reactive species. This phenomenon is often referred to as self-quenching. Kim et al. [39] observed that the degradation of micropollutants in the Fe(II)/PAA process at an initial pH of 3 increased from 69.7% to 100% over 60 min when the PAA concentration was raised from 50 μM to 500 μM, with [Fe(II)]0 fixed at 100 μM. However, further increasing the PAA concentration to 1000 μM reduced the degradation rate, thereby inhibiting micropollutant degradation. This decline can be attributed to the excessive PAA quenching of reactive radicals, which subsequently results in the production of less reactive species (Equations (23) and (24)) [39,57].
C H 3 C O O O H   + H O     C H 3 C ( O ) + H 2   O + O 2
C H 3 C O O O H   + H O     C H 3 C O O H   + H O 2
Kim et al. [20] also observed that in the Co(II)/PAA process, CBZ degradation significantly increased from 24.3% to 91.9% within 30 min as the PAA dosages were increased ([PAA]0 = 20, 50, 100, and 200 μM). This enhancement was reflected in the pseudo-first-order rate constant, which rose from (6.98 ± 0.04) × 10 4 to (4.83 ± 0.67) × 10 3   s 1 with increasing PAA concentrations. The observed trend can be attributed to the increased generation of RS in the solution at higher PAA concentrations. Zhang et al. [93] evaluated the effects of the initial PAA concentration on k o b s in the UV/PAA system for DCF degradation. They observed that k o b s increased gradually from 0.0967 to 0.182 min−1 as the PAA concentration was raised from 10 to 150 μM. However, k o b s increased proportionally with the rise in the PAA dose up to 50 μM, likely due to the enhanced production of C H 3 C O O O and H O radicals. Unlike the linear increase, k o b s continued to rise at higher PAA concentrations, but the rate of increase slowed. This behavior is likely attributed to the dual effect of excess PAA scavenging H O   radicals and competing with DCF for UV light absorption, thereby inhibiting direct photolysis.
When assessing the effects of PAA concentration on the removal of CECs in PAA-based systems, it is essential to maintain a constant pH value. Despite its importance, this detail is often overlooked in numerous research studies.

6.2. Effects of Catalyst Dosage

Degradation efficiency in AOPs can be more effectively enhanced by introducing a catalyst rather than simply increasing the oxidant concentration. Extensive research has demonstrated that higher catalyst dosages promote PAA activation, thereby improving the removal of CECs [39]. However, an excessively high catalyst concentration may lead to radical scavenging by the catalyst itself, ultimately reducing the degradation efficiency, which could either plateau or even decline. For example, Kim et al. [39] applied the Fe(II)/PAA process to the removal of CECs and investigated the effect of Fe(II) concentrations ranging from 20 to 500 μM. They observed a modest increase in the removal rate, from 63.1% to 75%, at concentrations between 20 and 100 μM. However, at 500 μM, the removal rate declined significantly to 46%. This finding suggests that excessively high Fe(II) concentrations can have a detrimental effect by depleting reactive radicals. A parallel trend was observed in the Co(II)/PAA process for CBZ degradation [20]. Increasing the concentration of Co(II) from 1 to 20 μM resulted in a linear increase in the degradation rate constant. However, when the Co(II) concentration exceeded 20 μM, the degradation rate was inhibited, likely due to radical scavenging by excess Co(II), which disrupted the degradation process’s efficiency.
On the other hand, for heterogeneous catalysts, increasing the catalyst dosage provides a greater number of available active catalytic sites, which significantly enhances the system’s ability to accelerate chemical reactions and improve overall efficiency. For example, Wang et al. [52] employed the C o F e 2 O 4 /PAA system with 100 μM of PAA to degrade SMX. In the absence of the catalyst, the SMX concentration showed negligible variation. However, increasing the C o F e 2 O 4 dose from 0.025 g·L−1 to 0.2 g·L−1 enhanced the SMX removal efficiency from 36.2% to 81.9% after 30 min, with k o b s increasing from 0.015 min–1 to 0.060 min–1. This improvement was attributed to the accelerated production of radicals via PAA decomposition at the active sites on the C o F e 2 O 4 surface, while the adsorption of SMX on the C o F e 2 O 4 surface remained minimal. An equivalent enhancement was also noted in the breakdown of Orange G (OG) by the C o 3 O 4 /PAA process [68]. The degradation efficiency of OG increased from 0 to 79% as the C o 3 O 4 dosage was increased from 0 to 0.01 g·L−1. Furthermore, even when the C o 3 O 4 loading was increased from 0.01 to 0.2 g·L−1, the degradation continued to improve.
Both homogeneous and heterogeneous catalyst types can exhibit comparable behavior; although raising the catalyst concentration generally accelerates reactions, there is a threshold beyond which it may result in unwanted side effects. For instance, Wang et al. [62] observed that as the Fe(II)–zeolite dosage increased from 0.2 to 0.8 g·L−1, the SMX removal rate constant rose from 0.013 to 0.13 min−1. In contrast, when the catalyst dosage was further increased to 1.0 g·L−1, the rate constant decreased to 0.77 min−1. Similarly, Zhou et al. [65] utilized a cobalt ferrite combined with a PAA system to degrade the organic contaminant rhodamine B (RhB). Increasing the C o F e 2 O 4 dosage from 0.3 to 0.5 g·L−1 enhanced the removal of RhB. However, when the dosage was further increased to 0.5–0.7 g·L−1, the degradation rate of RhB remained relatively stable. At a concentration of 1.0 g·L−1, the degradation rate was significantly inhibited, likely due to the agglomeration of magnetic nanoparticles.

6.3. Effects of pH

In PAA-based AOPs, pH plays a critical role, influencing the behavior of oxidants, catalysts, radical formation, and target contaminant interactions. A particularly important aspect is its impact on the acid–base equilibrium of PAA, which directly affects both its stability and reactivity. With a dissociation constant (pKa) of 8.2, the protonated form of PAA predominates at pH values below 8.2, whereas the deprotonated anionic form (PAA) becomes dominant at higher pH levels.
Regarding the UV/PAA process, PAA ( E P A A = 1.005 V) is a less potent oxidation agent than PAA [154]. Despite this, PAA exhibits a greater molar absorption coefficient at 254 nm compared to PAA, with values of 41.6 M−1cm−1 and 8.0 M−1cm−1, respectively, indicating that the photolysis of PAA and the generation of radicals occur more efficiently under alkaline conditions [155]. For instance, Zhang et al. [93] systematically studied the effects of pH on DCF degradation in the UV/PAA process and identified an optimal pH of 8.5. Within the investigated pH range of 3.0 to 8.5, the k o b s increased progressively from 0.096 to 0.1902 min−1, emphasizing the improved degradation efficiency under alkaline conditions. Nevertheless, operating under strongly alkaline conditions can compromise the effectiveness of the UV/PAA process, attributable to the decreased redox potential of the reactive H O , as described by the Nernst equation. This is further evidenced by Zhang et al. [93], who observed a reduction in the k o b s to 0.1536 min−1 at a pH of 11.0, compared to its value at a pH of 8.5.
Pertaining to the metal/PAA system, pH significantly influences the equilibrium states of both metal ions and PAA, thereby affecting their speciation and reactivity [156]. The interactions between transition metals and PAA depend on electron exchange, with the neutral form of PAA exhibiting a greater tendency to accept electrons compared to its deprotonated counterpart [157]. As mentioned earlier, PAA predominantly exists in its neutral form at acid to neutral pH levels. This aligns with studies on SMX degradation using the Co/PAA process, which demonstrate its pH dependence, with optimal performance observed at a near-neutral pH [77]. On the flip side, Liu et al. [41] demonstrated that under acidic conditions, the combination of Co(II) and PAA facilitates the generation of Co(IV) species, which are highly effective in degrading refractory organic contaminants, such as CBZ. In contrast, increasing the solution pH to neutral conditions, which are more commonly employed, appears to shift the mechanism, favoring the conversion of Co(IV) into oxygen-centered radicals.
Furthermore, throughout the process of heterogeneous PAA activation, the impact of pH varies depending on the type of catalyst employed, such as metal oxides. pH can significantly influence the surface charges of the metal oxide, which in turn affects their interaction with PAA. For instance, a neutral pH was found to be favorable for the degradation of OG in the C o 3 O 4 /PAA system due to the favorable electrostatic interaction between PAA and the catalyst surface. Under acidic conditions, the excess H + ions form hydrogen bonds with the O–O fragment of PAA, hindering its interaction with the positively charged C o 3 O 4 surface. Conversely, under alkaline conditions, electrostatic repulsion between the negatively charged oxide surface and PAA similarly reduces the catalyst’s effectiveness in interacting with PAA [68]. Similarly, metal–organic framework gels (MOGs) have emerged as highly effective catalysts for PAA activation, with performance is significantly influenced by pH. For example, Zheng et al. [158] investigated the FeMn bimetallic MOG with a hierarchical porous structure (FeMn13BTC)/PAA system for the degradation of ofloxacin (OFX). Their pH influence experiments revealed that the presence of H + ions hindered the system, indicating that OFX degradation was more effective under alkaline conditions. The isoelectric point of FeMn13BTC was determined to be at a pH of 2.81, meaning its surface becomes negatively charged at pH values above this threshold. However, OFX degradation under alkaline conditions was not suppressed by electrostatic repulsion between FeMn13BTC and PAA. Instead, the enhanced degradation observed in alkaline conditions was attributed to increased PAA hydrolysis at higher pH levels, resulting in more efficient PAA activation based on the FeMn13BTC/PAA system. Concerning carbon-based catalysts, the application of PAA under activated biochar (ABC) catalysis exhibits distinct mechanisms. In this context, ABC, derived from nut shells, was employed as an eco-friendly catalyst for the degradation of sulfamethazine (SMZ). Zhang et al. [103] demonstrated that the effective degradation of SMZ at an acidic pH (<4.62) was driven by the electrostatic attraction between the positively charged ABC surface and the dissociated acidic form of SMZ (SMZ). Furthermore, pH plays a crucial role in both the catalytic efficiency and adsorption behavior of carbon-based materials, thereby influencing their performance in the degradation process [103].
Understanding these pH-dependent behaviors is essential for choosing a suitable catalyst for environmental applications.

7. Challenges and Future Perspectives

Various strategies have been employed to enhance PAA activation for the efficient degradation of CECS. These processes can produce a diverse range of RS, including free radicals such as H O , C H 3 C ( O ) O , and C H 3 C ( O ) O O , as well as non-radical species like 1 O 2 and HVMs. Each of these RS contributes to the degradation of CECs to varying extents, depending on the specific activation pathway and reaction conditions. However, despite the promising development potential of PAA-based AOPs in recent years, several critical challenges persist and must be thoroughly addressed in future research to fully realize their effectiveness and applicability.
(1)
A comprehensive study of the selective oxidation mechanisms of both non-radical and organic free radical pathways is essential. Non-radical mechanisms, such as 1 O 2 and HVMs, have emerged as key areas of interest. Crucial subjects for further investigation involve the precise detection and quantification of C H 3 C ( O ) O O and C H 3 C ( O ) O , as well as exploring the oxidation capacity of organic radicals toward pollutants. Additionally, the differences in product accumulation and toxicity alterations between the free radical and non-free radical pathways for the same pollutant require further investigation.
(2)
Conducting parameter optimization techniques and their applications; creating analytical approaches for three-dimensional and other multidimensional factors is a key strategy for obtaining more sophisticated parameter optimization.
(3)
Investigating affordable and sustainable approaches for activating PAA, such as solar irradiation, holds significant potential for the degradation of CECs. The solar irradiation/PAA process offers advantages like easy accessibility, the absence of the need for additional chemicals, and its renewable nature, making it a promising avenue for further development and application.
(4)
Developing catalysts with enhanced catalytic efficiency: PAA-based AOPs primarily focus on metal-derived catalysts, with C o 2 + being the most commonly studied activator. While C o 2 + exhibits high catalytic performance at a neutral pH, its release into the environment poses a secondary source of pollution, necessitating careful consideration in future applications. Therefore, iron activation is considered a more environmentally friendly option. However, iron-based systems are highly pH-sensitive and involve additional chemicals for pH regulation. Moreover, although heterogeneous catalysts release metal ions at lower concentrations, these ions can still accumulate to levels exceeding water treatment standards. To mitigate these issues, encapsulating metal nanoparticles within a carbon or polymer layer offers a promising solution. This encapsulation enables the precise tuning of the electronic structure and work function, while leveraging the electron tunneling effect at the composite interface to optimize catalytic performance. Such advancements could enhance the efficiency and environmental compatibility of PAA-based AOPs, paving the way for more sustainable water treatment technologies.
(5)
Given that most current studies use ultrapure water matrices, and few explore real wastewater conditions, the influence of coexisting ions and natural organic matter remains poorly understood. Further research using real water matrices is critically needed to better predict process performance under practical conditions.
(6)
Additionally, while the laboratory-scale efficiency of PAA-based AOPs has been demonstrated, few studies have addressed the development of scalable reactor designs or process integration strategies for industrial applications. Future work should focus on conducting pilot-scale studies to evaluate the use of PAA-based AOPs in large-scale systems, focusing on managing multiple pollutants in real-world water treatment processes.
(7)
Additionally, efforts should be made to integrate PAA-based AOPs with technologies such as electrochemistry, membrane filtration, and biological treatments to improve overall efficiency.

Author Contributions

Conceptualization, B.B., S.C.E. and M.Á.S.; resources, M.Á.S.; writing—original draft preparation, B.B. and S.C.E.; writing—review and editing, S.C.E.; visualization, B.B. and M.Á.S.; supervision, M.Á.S.; project administration, M.Á.S.; funding acquisition, M.Á.S. All authors have read and agreed to the published version of the manuscript.

Funding

The authors acknowledge financial support from this research that has been financially supported by Xunta de Galicia (ED431C 2021-43).

Data Availability Statement

All data are contained within the article.

Conflicts of Interest

The authors declare the following financial interests/personal relationships which may be considered as potential competing interests: M.A. Sanromán reports financial support was provided by Xunta de Galicia.

References

  1. Pal, A.; He, Y.; Jekel, M.; Reinhard, M.; Gin, K.Y.-H. Emerging Contaminants of Public Health Significance as Water Quality Indicator Compounds in the Urban Water Cycle. Environ. Int. 2014, 71, 46–62. [Google Scholar] [CrossRef] [PubMed]
  2. Li, P.; Wu, J. Drinking Water Quality and Public Health. Expo. Health 2019, 11, 73–79. [Google Scholar] [CrossRef]
  3. Kotowska, U.; Karpińska, J.; Kiejza, D.; Ratkiewicz, A.; Piekutin, J.; Makarova, K.; Olchowik-Grabarek, E. Oxidation of Contaminants of Emerging Concern by Combination of Peracetic Acid with Iron Ions and Various Types of Light Radiation–Optimization, Kinetics, Removal Efficiency and Mechanism Investigation. J. Mol. Liq. 2023, 369, 120859. [Google Scholar] [CrossRef]
  4. Richardson, S.D.; Kimura, S.Y. Water Analysis: Emerging Contaminants and Current Issues. Anal. Chem. 2016, 88, 546–582. [Google Scholar] [CrossRef]
  5. Wu, X.; Ren, J.; Xu, Q.; Xiao, Y.; Li, X.; Peng, Y. Priority Screening of Contaminant of Emerging Concern (CECs) in Surface Water from Drinking Water Sources in the Lower Reaches of the Yangtze River Based on Exposure-Activity Ratios (EARs). Sci. Total Environ. 2023, 856, 159016. [Google Scholar] [CrossRef]
  6. Ahmad, A.; Kurniawan, S.B.; Abdullah, S.R.S.; Othman, A.R.; Hasan, H.A. Contaminants of Emerging Concern (CECs) in Aquaculture Effluent: Insight into Breeding and Rearing Activities, Alarming Impacts, Regulations, Performance of Wastewater Treatment Unit and Future Approaches. Chemosphere 2022, 290, 133319. [Google Scholar] [CrossRef] [PubMed]
  7. Kasonga, T.K.; Coetzee, M.A.A.; Kamika, I.; Ngole-Jeme, V.M.; Momba, M.N.B. Endocrine-Disruptive Chemicals as Contaminants of Emerging Concern in Wastewater and Surface Water: A Review. J. Environ. Manag. 2021, 277, 111485. [Google Scholar] [CrossRef]
  8. Gavrilescu, M. Environmental Biotechnology: Achievements, Opportunities and Challenges. Dyn. Biochem. Process Biotechnol. Mol. Biol. 2010, 4, 1–36. [Google Scholar]
  9. Kumari, P.; Kumar, A. Advanced Oxidation Process: A Remediation Technique for Organic and Non-Biodegradable Pollutant. Results Surf. Interfaces 2023, 11, 100122. [Google Scholar] [CrossRef]
  10. Ahmed, S.F.; Mofijur, M.; Nuzhat, S.; Chowdhury, A.T.; Rafa, N.; Uddin, M.A.; Inayat, A.; Mahlia, T.M.I.; Ong, H.C.; Chia, W.Y. Recent Developments in Physical, Biological, Chemical, and Hybrid Treatment Techniques for Removing Emerging Contaminants from Wastewater. J. Hazard. Mater. 2021, 416, 125912. [Google Scholar] [CrossRef]
  11. Khan, J.A.; Sayed, M.; Khan, S.; Shah, N.S.; Dionysiou, D.D.; Boczkaj, G. Advanced Oxidation Processes for the Treatment of Contaminants of Emerging Concern. In Contaminants of Emerging Concern in Water and Wastewater; Elsevier: Amsterdam, The Netherlands, 2020; pp. 299–365. [Google Scholar]
  12. Deng, Y.; Zhao, R. Advanced Oxidation Processes (AOPs) in Wastewater Treatment. Curr. Pollut. Rep. 2015, 1, 167–176. [Google Scholar] [CrossRef]
  13. Giannakis, S.; Lin, K.-Y.A.; Ghanbari, F. A Review of the Recent Advances on the Treatment of Industrial Wastewaters by Sulfate Radical-Based Advanced Oxidation Processes (SR-AOPs). Chem. Eng. J. 2021, 406, 127083. [Google Scholar] [CrossRef]
  14. Scaria, J.; Nidheesh, P.V. Comparison of Hydroxyl-Radical-Based Advanced Oxidation Processes with Sulfate Radical-Based Advanced Oxidation Processes. Curr. Opin. Chem. Eng. 2022, 36, 100830. [Google Scholar] [CrossRef]
  15. Liu, Y.; Liu, L.; Wang, Y. A Critical Review on Removal of Gaseous Pollutants Using Sulfate Radical-Based Advanced Oxidation Technologies. Environ. Sci. Technol. 2021, 55, 9691–9710. [Google Scholar] [CrossRef] [PubMed]
  16. Liu, Y.; He, X.; Fu, Y.; Dionysiou, D.D. Degradation Kinetics and Mechanism of Oxytetracycline by Hydroxyl Radical-Based Advanced Oxidation Processes. Chem. Eng. J. 2016, 284, 1317–1327. [Google Scholar] [CrossRef]
  17. Li, J.; Yang, T.; Zeng, G.; An, L.; Jiang, J.; Ao, Z.; Ma, J. Ozone-and Hydroxyl Radical-Induced Degradation of Micropollutants in a Novel UVA-LED-Activated Periodate Advanced Oxidation Process. Environ. Sci. Technol. 2023, 57, 18607–18616. [Google Scholar] [CrossRef] [PubMed]
  18. Liou, S.-Y.; Dodd, M.C. Evaluation of Hydroxyl Radical and Reactive Chlorine Species Generation from the Superoxide/Hypochlorous Acid Reaction as the Basis for a Novel Advanced Oxidation Process. Water Res. 2021, 200, 117142. [Google Scholar] [CrossRef]
  19. Ao, X.; Eloranta, J.; Huang, C.-H.; Santoro, D.; Sun, W.; Lu, Z.; Li, C. Peracetic Acid-Based Advanced Oxidation Processes for Decontamination and Disinfection of Water: A Review. Water Res. 2021, 188, 116479. [Google Scholar] [CrossRef]
  20. 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]
  21. 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]
  22. Wojnárovits, L.; Homlok, R.; Kovács, K.; Tegze, A.; Takács, E. Oxidation and Mineralization Rates of Harmful Organic Chemicals in Hydroxyl Radical Induced Reactions. Ecotoxicol. Environ. Saf. 2024, 281, 116669. [Google Scholar] [CrossRef] [PubMed]
  23. Krystynik, P. Advanced Oxidation Processes (AOPs)—Utilization of Hydroxyl Radical and Singlet Oxygen. In Reactive Oxygen Species; IntechOpen: London, UK, 2021. [Google Scholar]
  24. Liu, Z.; Demeestere, K.; Van Hulle, S. Comparison and Performance Assessment of Ozone-Based AOPs in View of Trace Organic Contaminants Abatement in Water and Wastewater: A Review. J. Environ. Chem. Eng. 2021, 9, 105599. [Google Scholar] [CrossRef]
  25. Zhou, Z.; Liu, X.; Sun, K.; Lin, C.; Ma, J.; He, M.; Ouyang, W. Persulfate-Based Advanced Oxidation Processes (AOPs) for Organic-Contaminated Soil Remediation: A Review. Chem. Eng. J. 2019, 372, 836–851. [Google Scholar] [CrossRef]
  26. Ghanbari, F.; Moradi, M. Application of Peroxymonosulfate and Its Activation Methods for Degradation of Environmental Organic Pollutants. Chem. Eng. J. 2017, 310, 41–62. [Google Scholar] [CrossRef]
  27. Li, Y.; Liu, Y.; Zhan, Y.; Zhang, Y.; Zhao, X.; Yang, M.; Ruan, W.; Zhang, Z.; Liang, X.; Ma, J. Peracetic Acid-Induced Nanoengineering of Fe-Based Metallic Glass Ribbon in Application of Efficient Drinking Water Treatment. Appl. Catal. B Environ. Energy 2024, 355, 124161. [Google Scholar] [CrossRef]
  28. Qiao, C.; Jia, W.; Tang, J.; Chen, C.; Wu, Y.; Liang, Y.; Du, J.; Wu, Q.; Feng, X.; Wang, H. Advances of Carbon-Based Materials for Activating Peracetic Acid in Advanced Oxidation Processes: A Review. Environ. Res. 2024, 263, 120058. [Google Scholar] [CrossRef]
  29. Shen, W.; Yang, L.; Zhou, Z.; Gao, H.; Zhou, X.; Zhang, Y.; Chen, J. Activation of Peracetic Acid by Thermally Modified Carbon Nanotubes: Organic Radicals Contribution and Active Sites Identification. Chem. Eng. J. 2023, 474, 145521. [Google Scholar] [CrossRef]
  30. Zhou, G.; Zhou, R.; Liu, Y.; Zhang, L.; Zhang, L.; Fu, Y. Efficient Degradation of Sulfamethoxazole Using Peracetic Acid Activated by Zero-Valent Cobalt. J. Environ. Chem. Eng. 2022, 10, 107783. [Google Scholar] [CrossRef]
  31. Wang, Z.; Chen, Z.; Li, Q.; Wang, J.; Cao, L.; Cheng, Y.; Yu, S.; Liu, Z.; Chen, Y.; Yue, S. Non-Radical Activation of Peracetic Acid by Powdered Activated Carbon for the Degradation of Sulfamethoxazole. Environ. Sci. Technol. 2023, 57, 10478–10488. [Google Scholar] [CrossRef]
  32. 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]
  33. 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]
  34. Chen, J.; Xu, J.; Liu, T.; Wang, Q.; Li, N.; Zhang, Y.; Yang, L.; Zhou, X. Oxidation of Tetracycline Antibiotics by Peracetic Acid: Reaction Kinetics, Mechanism, and Antibacterial Activity Change. Chem. Eng. J. 2022, 431, 134190. [Google Scholar] [CrossRef]
  35. Kim, J.; Huang, C.-H. Reactivity of Peracetic Acid with Organic Compounds: A Critical Review. ACS EST Water 2020, 1, 15–33. [Google Scholar] [CrossRef]
  36. 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]
  37. Liu, F.; Zou, Y.; Liang, H.; Hu, J.; Li, Y.; Lin, L.; Li, X.; Li, B. Trace Co (II) Triggers Peracetic Acid Activation in Phosphate Buffer: New Insights into the Oxidative Species Responsible for Ciprofloxacin Removal. J. Hazard. Mater. 2024, 467, 133638. [Google Scholar] [CrossRef] [PubMed]
  38. Caretti, C.; Lubello, C. Wastewater Disinfection with PAA and UV Combined Treatment: A Pilot Plant Study. Water Res. 2003, 37, 2365–2371. [Google Scholar] [CrossRef] [PubMed]
  39. 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]
  40. 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] [PubMed]
  41. Liu, B.; Guo, W.; Jia, W.; Wang, H.; Zheng, S.; Si, Q.; Zhao, Q.; Luo, H.; Jiang, J.; Ren, N. Insights into the Oxidation of Organic Contaminants by Co (II) Activated Peracetic Acid: The Overlooked Role of High-Valent Cobalt-Oxo Species. Water Res. 2021, 201, 117313. [Google Scholar] [CrossRef]
  42. 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 EST Eng. 2021, 1, 1432–1440. [Google Scholar] [CrossRef]
  43. 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]
  44. Wu, J.; Zou, J.; Lin, J.; Li, S.; He, L.; Wu, Z.; Li, Q.; Gong, C.; Ma, J. Overlooked Role of Coexistent Hydrogen Peroxide in Activated Peracetic Acid by Cu (II) for Enhanced Oxidation of Organic Contaminants. Environ. Sci. Technol. 2024, 58, 15741–15754. [Google Scholar] [CrossRef] [PubMed]
  45. Li, R.; Manoli, K.; Kim, J.; Feng, M.; Huang, C.-H.; Sharma, V.K. Peracetic Acid–Ruthenium (III) Oxidation Process for the Degradation of Micropollutants in Water. Environ. Sci. Technol. 2021, 55, 9150–9160. [Google Scholar] [CrossRef] [PubMed]
  46. Luukkonen, T.; Pehkonen, S.O. Peracids in Water Treatment: A Critical Review. Crit. Rev. Environ. Sci. Technol. 2017, 47, 1–39. [Google Scholar] [CrossRef]
  47. Bell, J.; Wen, Y.; Ma, X.; McDonald, T.J.; Huang, C.-H.; Sharma, V.K. Interaction of Peracetic Acid with Chromium (III): Understanding Degradation of Coexisting Organic Pollutants in Water. J. Hazard. Mater. 2022, 438, 129537. [Google Scholar] [CrossRef]
  48. Wang, J.; Wang, Z.; Cheng, Y.; Cao, L.; Xie, P.; Ma, J. Molybdenum Disulfide (MoS2) Promoted Sulfamethoxazole Degradation in the Fe (III)/Peracetic Acid Process. Sep. Purif. Technol. 2022, 281, 119854. [Google Scholar] [CrossRef]
  49. 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]
  50. 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]
  51. Yang, K.; Zhai, Z.; Liu, H.; Zhao, T.; Yuan, D.; Jiao, T.; Zhang, Q.; Tang, S. Peracetic Acid Activation by Natural Chalcopyrite for Metronidazole Degradation: Unveiling the Effects of Cu-Fe Bimetallic Sites and Sulfur Species. Sep. Purif. Technol. 2023, 305, 122500. [Google Scholar] [CrossRef]
  52. 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 2021, 280, 119422. [Google Scholar] [CrossRef]
  53. 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] [PubMed]
  54. Sharma, S.; Mukhopadhyay, M.; Murthy, Z.V.P. UV/Peroxyacetic Acid Mediated Chlorophenol Congener Degradation. CLEAN–Soil Air Water 2014, 42, 276–283. [Google Scholar] [CrossRef]
  55. 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]
  56. Yao, K.; Fang, L.; Liao, P.; Chen, H. Ultrasound-Activated Peracetic Acid to Degrade Tetracycline Hydrochloride: Efficiency and Mechanism. Sep. Purif. Technol. 2023, 306, 122635. [Google Scholar] [CrossRef]
  57. Yuan, D.; Yang, K.; Pan, S.; Xiang, Y.; Tang, S.; Huang, L.; Sun, M.; Zhang, X.; Jiao, T.; Zhang, Q. Peracetic Acid Enhanced Electrochemical Advanced Oxidation for Organic Pollutant Elimination. Sep. Purif. Technol. 2021, 276, 119317. [Google Scholar] [CrossRef]
  58. Dai, Y.; Qi, C.; Cao, H.; Wen, Y.; Zhao, Y.; Xu, C.; Yang, S.; He, H. Enhanced Degradation of Sulfamethoxazole by Microwave-Activated Peracetic Acid under Alkaline Condition: Influencing Factors and Mechanism. Sep. Purif. Technol. 2022, 288, 120716. [Google Scholar] [CrossRef]
  59. Dai, C.; Li, S.; Duan, Y.; Leong, K.H.; Liu, S.; Zhang, Y.; Zhou, L.; Tu, Y. Mechanisms and Product Toxicity of Activated Carbon/Peracetic Acid for Degradation of Sulfamethoxazole: Implications for Groundwater Remediation. Water Res. 2022, 216, 118347. [Google Scholar] [CrossRef] [PubMed]
  60. Wu, L.; Li, Z.; Cheng, P.; She, Y.; Wang, W.; Tian, Y.; Ma, J.; Sun, Z. Efficient Activation of Peracetic Acid by Mixed Sludge Derived Biochar: Critical Role of Persistent Free Radicals. Water Res. 2022, 223, 119013. [Google Scholar] [CrossRef]
  61. Correa-Sanchez, S.; Peñuela, G.A. Peracetic Acid-Based Advanced Oxidation Processes for the Degradation of Emerging Pollutants: A Critical Review. J. Water Process Eng. 2022, 49, 102986. [Google Scholar] [CrossRef]
  62. 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]
  63. Yang, T.; An, L.; Zeng, G.; Jiang, M.; Li, J.; Liu, C.; Jia, J.; Ma, J. Efficient Removal of P-Arsanilic Acid and Arsenite by Fe (II)/Peracetic Acid (Fe (II)/PAA) and PAA Processes. Water Res. 2023, 241, 120091. [Google Scholar] [CrossRef] [PubMed]
  64. 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]
  65. Zhou, G.; Fu, Y.; Zhou, R.; Zhang, L.; Zhang, L.; Deng, J.; Liu, Y. Efficient Degradation of Organic Contaminants by Magnetic Cobalt Ferrite Combined with Peracetic Acid. Process Saf. Environ. Prot. 2022, 160, 376–384. [Google Scholar] [CrossRef]
  66. Zhang, P.; Zhang, X.; Zhao, X.; Jing, G.; Zhou, Z. Activation of Peracetic Acid with Zero-Valent Iron for Tetracycline Abatement: The Role of Fe (II) Complexation with Tetracycline. J. Hazard. Mater. 2022, 424, 127653. [Google Scholar] [CrossRef]
  67. 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]
  68. 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]
  69. Zhou, X.; Wu, H.; Zhang, L.; Liang, B.; Sun, X.; Chen, J. Activation of Peracetic Acid with Lanthanum Cobaltite Perovskite for Sulfamethoxazole Degradation under a Neutral PH: The Contribution of Organic Radicals. Molecules 2020, 25, 2725. [Google Scholar] [CrossRef]
  70. Zhou, R.; Zhou, G.; Liu, Y.; Liu, S.; Wang, S.; Fu, Y. Activated Peracetic Acid by Mn3O4 for Sulfamethoxazole Degradation: A Novel Heterogeneous Advanced Oxidation Process. Chemosphere 2022, 306, 135506. [Google Scholar] [CrossRef] [PubMed]
  71. Xing, D.; Shao, S.; Yang, Y.; Zhou, Z.; Jing, G.; Zhao, X. Mechanistic Insights into the Efficient Activation of Peracetic Acid by Pyrite for the Tetracycline Abatement. Water Res. 2022, 222, 118930. [Google Scholar] [CrossRef]
  72. Cheng, Y.; Wang, Z.; Wang, J.; Cao, L.; Chen, Z.; Chen, Y.; Liu, Z.; Xie, P.; Ma, J. New Insights into the Degradation of Micro-Pollutants in the Hydroxylamine Enhanced Fe (II)/Peracetic Acid Process: Contribution of Reactive Species and Effects of PH. J. Hazard. Mater. 2023, 441, 129885. [Google Scholar] [CrossRef]
  73. Yang, S.; Liang, Z.-H.; Wen, Y.; He, C.-S.; Xiong, Z.; Du, Y.; Liu, Y.; Zhang, H.; Zhou, P.; Mu, Y. Gallic Acid Accelerates the Oxidation Ability of the Peracetic Acid/Fe (III) System for Bisphenol A Removal: Fate of Various Radicals. ACS EST Eng. 2023, 3, 271–282. [Google Scholar] [CrossRef]
  74. Lin, J.; Hu, Y.; Xiao, J.; Huang, Y.; Wang, M.; Yang, H.; Zou, J.; Yuan, B.; Ma, J. Enhanced Diclofenac Elimination in Fe (II)/Peracetic Acid Process by Promoting Fe (III)/Fe (II) Cycle with ABTS as Electron Shuttle. Chem. Eng. J. 2021, 420, 129692. [Google Scholar] [CrossRef]
  75. Li, C.; Yuan, D.; Yang, K.; Wang, H.; Wang, Z.; Zhang, Q.; Tang, S. Reutilization of Pyrite Tailings in Peracetic Acid-Based Advanced Oxidation Process for Water Purification. Sep. Purif. Technol. 2025, 354, 129155. [Google Scholar] [CrossRef]
  76. Yao, M.; Zhang, S.; Xie, M.; Zhao, L.; Zhao, R.-S. Efficient Activation of Peracetic Acid by Cobalt Modified Nitrogen-Doped Carbon Nanotubes for Drugs Degradation: Performance and Mechanism Insight. Chemosphere 2024, 358, 142277. [Google Scholar] [CrossRef]
  77. 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. 2019, 54, 464–475. [Google Scholar] [CrossRef] [PubMed]
  78. Haghshenas, D.F.; Darvishi, D.; Moshefi, S.Z.; Keshavarz, A.E.; Sadrnezhad, S.K. Leaching Recovery of Zinc, Cobalt and Manganese from Zinc Purification Residue. Int. J. Eng. 2007, 20, 133–140. [Google Scholar]
  79. Moore, T.A.; Black, A.; Centeno, J.A.; Harding, J.S.; Trumm, D.A. Metal Contaminants in New Zealand. Sources, Treatments, and Effects on Ecology and Human Health; Resolutionz Press: Sydney, NSW, Australia, 2005. [Google Scholar]
  80. Barceloux, D.G.; Barceloux, D. Cobalt. J. Toxicol. Clin. Toxicol. 1999, 37, 201–216. [Google Scholar] [CrossRef]
  81. Zhou, R.; Zhou, G.; Liu, Y.; Wang, S.; Fu, Y. Cobalt Doped Graphitic Carbon Nitride as an Effective Catalyst for Peracetic Acid to Degrade Sulfamethoxazole. RSC Adv. 2022, 12, 13810–13819. [Google Scholar] [CrossRef]
  82. Zhang, Z.-Q.; Duan, P.-J.; Zheng, J.-X.; Xie, Y.-Q.; Bai, C.-W.; Sun, Y.-J.; Chen, X.-J.; Chen, F.; Yu, H.-Q. Nano-Island-Encapsulated Cobalt Single-Atom Catalysts for Breaking Activity-Stability Trade-off in Fenton-like Reactions. Nat. Commun. 2025, 16, 115. [Google Scholar] [CrossRef]
  83. Lv, K.; Ling, L.; Lu, Q.; Lu, J.; Zhou, Y.; Zhou, Y. Reducing Cobalt Leaching in Disinfectant Degradation: Insights from Chitosan-Derived Carbon Encapsulated Co9S8. Sep. Purif. Technol. 2024, 344, 127207. [Google Scholar] [CrossRef]
  84. Taujale, S.; Baratta, L.R.; Huang, J.; Zhang, H. Interactions in Ternary Mixtures of MnO2, Al2O3, and Natural Organic Matter (NOM) and the Impact on MnO2 Oxidative Reactivity. Environ. Sci. Technol. 2016, 50, 2345–2353. [Google Scholar] [CrossRef] [PubMed]
  85. Babuponnusami, A.; Muthukumar, K. A Review on Fenton and Improvements to the Fenton Process for Wastewater Treatment. J. Environ. Chem. Eng. 2014, 2, 557–572. [Google Scholar] [CrossRef]
  86. Zhao, W.; Duan, Z.; Zheng, Z.; Li, B. Efficient Diclofenac Removal by Superoxide Radical and Singlet Oxygen Generated in Surface Mn (II)/(III)/(IV) Cycle Dominated Peroxymonosulfate Activation System: Mechanism and Product Toxicity. Chem. Eng. J. 2022, 433, 133742. [Google Scholar] [CrossRef]
  87. Ember, E.; Rothbart, S.; Puchta, R.; van Eldik, R. Metal Ion-Catalyzed Oxidative Degradation of Orange II by H2O2. High Catalytic Activity of Simple Manganese Salts. New J. Chem. 2009, 33, 34–49. [Google Scholar] [CrossRef]
  88. Zong, Y.; Shao, Y.; Ji, W.; Zeng, Y.; Xu, J.; Liu, W.; Xu, L.; Wu, D. Trace Mn (II)-Catalyzed Periodate Oxidation of Organic Contaminants Not Relying on Any Transient Reactive Species: The Substrate-Dependent Dual Roles of in-Situ Formed Colloidal MnO2. Chem. Eng. J. 2023, 451, 139106. [Google Scholar] [CrossRef]
  89. Kim, J.; Wang, J.; Ashley, D.C.; Sharma, V.K.; Huang, C.-H. Picolinic Acid-Mediated Catalysis of Mn (II) for Peracetic Acid Oxidation Processes: Formation of High-Valent Mn Species. Environ. Sci. Technol. 2023, 57, 18929–18939. [Google Scholar] [CrossRef]
  90. Dong, J.; Dong, H.; Xiao, J.; Li, L.; Huang, D.; Zhao, M. Enhanced Degradation of Micropollutants in a Peracetic Acid/Mn (II) System with EDDS: An Investigation of the Role of Mn Species. Environ. Sci. Technol. 2024, 58, 12179–12188. [Google Scholar] [CrossRef]
  91. Liu, Y.; Zhou, R.; Tang, Y.; Li, X.; Xu, L.; Fu, Y. Enhanced Mn (II)/Peracetic Acid by Nitrilotriacetic Acid to Degrade Organic Contaminants: Role of Mn (V) and Organic Radicals. Sci. Rep. 2024, 14, 29686. [Google Scholar] [CrossRef]
  92. 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]
  93. 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]
  94. Paul, A.; Dziallas, C.; Zwirnmann, E.; Gjessing, E.T.; Grossart, H.-P. UV Irradiation of Natural Organic Matter (NOM): Impact on Organic Carbon and Bacteria. Aquat. Sci. 2012, 74, 443–454. [Google Scholar] [CrossRef]
  95. Sharma, S.; Mukhopadhyay, M.; Murthy, Z.V.P. Degradation of 4-Chlorophenol in Wastewater by Organic Oxidants. Ind. Eng. Chem. Res. 2010, 49, 3094–3098. [Google Scholar] [CrossRef]
  96. 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]
  97. Miao, F.; Yue, X.; Cheng, C.; Chen, X.; Ren, W.; Zhang, H. Insights into the Mechanism of Carbocatalysis for Peracetic Acid Activation: Kinetic Discernment and Active Site Identification. Water Res. 2022, 227, 119346. [Google Scholar] [CrossRef]
  98. Kong, D.; Zhao, Y.; Fan, X.; Wang, X.; Li, J.; Wang, X.; Nan, J.; Ma, J. Reduced Graphene Oxide Triggers Peracetic Acid Activation for Robust Removal of Micropollutants: The Role of Electron Transfer. Environ. Sci. Technol. 2022, 56, 11707–11717. [Google Scholar] [CrossRef] [PubMed]
  99. 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]
  100. Tan, H.; Li, J.; He, M.; Li, J.; Zhi, D.; Qin, F.; Zhang, C. Global Evolution of Research on Green Energy and Environmental Technologies: A Bibliometric Study. J. Environ. Manag. 2021, 297, 113382. [Google Scholar] [CrossRef]
  101. Huong, P.T.; Jitae, K.; Al Tahtamouni, T.M.; Tri, N.L.M.; Kim, H.-H.; Cho, K.H.; Lee, C. Novel Activation of Peroxymonosulfate by Biochar Derived from Rice Husk toward Oxidation of Organic Contaminants in Wastewater. J. Water Process Eng. 2020, 33, 101037. [Google Scholar] [CrossRef]
  102. Hou, J.; Xu, J.; Tang, R.; Min, Y.; Eitssayeam, S.; Shi, P. Activation of Peracetic Acid by Biochar for Sulfamethoxazole Degradation: Revealing Roles of the Active Sites. J. Taiwan. Inst. Chem. Eng. 2023, 152, 105184. [Google Scholar] [CrossRef]
  103. Zhang, Z.; Duan, Y.; Dai, C.; Li, S.; Chen, Y.; Tu, Y.; Leong, K.H.; Zhou, L. Oxidation of Sulfamethazine by Peracetic Acid Activated with Biochar: Reactive Oxygen Species Contribution and Toxicity Change. Environ. Pollut. 2022, 313, 120170. [Google Scholar] [CrossRef]
  104. Park, M.; Kim, N.; Lee, J.; Gu, M.; Kim, B.-S. Versatile Graphene Oxide Nanosheets via Covalent Functionalization and Their Applications. Mater. Chem. Front. 2021, 5, 4424–4444. [Google Scholar] [CrossRef]
  105. Shahzad, A.; Jawad, A.; Ifthikar, J.; Chen, Z.; Chen, Z. The Hetero-Assembly of Reduced Graphene Oxide and Hydroxide Nanosheets as Superlattice Materials in PMS Activation. Carbon 2019, 155, 740–755. [Google Scholar] [CrossRef]
  106. Jothinathan, L.; Hu, J. Kinetic Evaluation of Graphene Oxide Based Heterogenous Catalytic Ozonation for the Removal of Ibuprofen. Water Res. 2018, 134, 63–73. [Google Scholar] [CrossRef]
  107. Farajollahi, A.; Marjani, A.P.; Pesyan, N.N.; Alamgholiloo, H. Efficient Degradation of Crystal Violet by GO/CuMn2O4 Nanocomposite via Peroxymonosulfate Activation. Appl. Surf. Sci. 2023, 622, 156903. [Google Scholar] [CrossRef]
  108. Tshangana, C.; Mubiayi, M.P.; Kuvarega, A.; Mamba, B.; Muleja, A. Advanced Oxidation Processes for Pharmaceutical Degradation and Disinfection of Wastewater: Peracetic Acid and Graphene Oxide Quantum Dots. Int. J. Environ. Sci. Technol. 2023, 20, 11997–12014. [Google Scholar] [CrossRef]
  109. Choi, J.; Tu, N.D.K.; Lee, S.-S.; Lee, H.; Kim, J.S.; Kim, H. Controlled Oxidation Level of Reduced Graphene Oxides and Its Effect on Thermoelectric Properties. Macromol. Res. 2014, 22, 1104–1108. [Google Scholar] [CrossRef]
  110. Olmez-Hanci, T.; Arslan-Alaton, I.; Gurmen, S.; Gafarli, I.; Khoei, S.; Safaltin, S.; Ozcelik, D.Y. Oxidative Degradation of Bisphenol A by Carbocatalytic Activation of Persulfate and Peroxymonosulfate with Reduced Graphene Oxide. J. Hazard. Mater. 2018, 360, 141–149. [Google Scholar] [CrossRef]
  111. Kong, D.; Zhao, Y.; Guo, H.; Han, M.; Fan, X.; Li, J.; He, X.; Ma, J. Unveiling the Direct Electron Transfer Regime of Peracetic Acid Activation: Quantitative Structure–Activity Relationship Analysis of Carbon Nanotube Catalysis. ACS EST Eng. 2023, 3, 1030–1041. [Google Scholar] [CrossRef]
  112. Tian, X.; Liu, S.; Zhang, B.; Wang, S.; Dong, S.; Liu, Y.; Feng, L.; Zhang, L. Carbonized Polyaniline-Activated Peracetic Acid Advanced Oxidation Process for Organic Removal: Efficiency and Mechanisms. Environ. Res. 2023, 219, 115035. [Google Scholar] [CrossRef]
  113. Li, D.; Duan, X.; Sun, H.; Kang, J.; Zhang, H.; Tade, M.O.; Wang, S. Facile Synthesis of Nitrogen-Doped Graphene via Low-Temperature Pyrolysis: The Effects of Precursors and Annealing Ambience on Metal-Free Catalytic Oxidation. Carbon 2017, 115, 649–658. [Google Scholar] [CrossRef]
  114. Chen, F.; Liu, L.; Wu, J.; Rui, X.; Chen, J.; Yu, Y. Single-atom Iron Anchored Tubular G-C3N4 Catalysts for Ultrafast Fenton-like Reaction: Roles of High-valency Iron-oxo Species and Organic Radicals. Adv. Mater. 2022, 34, 2202891. [Google Scholar] [CrossRef] [PubMed]
  115. Li, M.; Sun, J.; Mei, Q.; Wei, B.; An, Z.; Cao, H.; Zhang, C.; Xie, J.; Zhan, J.; Wang, W. Acetaminophen Degradation by Hydroxyl and Organic Radicals in the Peracetic Acid-Based Advanced Oxidation Processes: Theoretical Calculation and Toxicity Assessment. J. Hazard. Mater. 2021, 416, 126250. [Google Scholar] [CrossRef]
  116. Mortensen, A. Scavenging of Acetylperoxyl Radicals and Quenching of Triplet Diacetyl by β-Carotene: Mechanisms and Kinetics. J. Photochem. Photobiol. B 2001, 61, 62–67. [Google Scholar] [CrossRef] [PubMed]
  117. Zhang, L.; Chen, J.; Zhang, Y.; Xu, Y.; Zheng, T.; Zhou, X. Highly Efficient Activation of Peracetic Acid by Nano-CuO for Carbamazepine Degradation in Wastewater: The Significant Role of H2O2 and Evidence of Acetylperoxy Radical Contribution. Water Res. 2022, 216, 118322. [Google Scholar] [CrossRef] [PubMed]
  118. Zhou, Y.; Ma, Z.; Tang, J.; Yan, N.; Du, Y.; Xi, S.; Wang, K.; Zhang, W.; Wen, H.; Wang, J. Immediate Hydroxylation of Arenes to Phenols via V-Containing All-Silica ZSM-22 Zeolite Triggered Non-Radical Mechanism. Nat. Commun. 2018, 9, 2931. [Google Scholar] [CrossRef]
  119. 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 2021, 297, 120475. [Google Scholar] [CrossRef]
  120. Zhang, L.; Chen, J.; Zheng, T.; Xu, Y.; Liu, T.; Yin, W.; Zhang, Y.; Zhou, X. Co–Mn Spinel Oxides Trigger Peracetic Acid Activation for Ultrafast Degradation of Sulfonamide Antibiotics: Unveiling Critical Role of Mn Species in Boosting Co Activity. Water Res. 2023, 229, 119462. [Google Scholar] [CrossRef]
  121. Chen, L.; Duan, J.; Du, P.; Sun, W.; Lai, B.; Liu, W. Accurate Identification of Radicals by In-Situ Electron Paramagnetic Resonance in Ultraviolet-Based Homogenous Advanced Oxidation Processes. Water Res. 2022, 221, 118747. [Google Scholar] [CrossRef]
  122. Lee, J.; Von Gunten, U.; Kim, J.-H. Persulfate-Based Advanced Oxidation: Critical Assessment of Opportunities and Roadblocks. Environ. Sci. Technol. 2020, 54, 3064–3081. [Google Scholar] [CrossRef]
  123. Wang, X.; Si, D.; Li, Y.; Chen, N.; Fang, G.; Zhu, C.; Zhou, D. Alcohols Radicals Can Efficiently Reduce Recalcitrant Perfluorooctanoic Acid. Water Res. 2023, 245, 120557. [Google Scholar] [CrossRef]
  124. Wang, X.; Qin, Y.; Zhu, L.; Tang, H. Nitrogen-Doped Reduced Graphene Oxide as a Bifunctional Material for Removing Bisphenols: Synergistic Effect between Adsorption and Catalysis. Environ. Sci. Technol. 2015, 49, 6855–6864. [Google Scholar] [CrossRef] [PubMed]
  125. Xu, M.; Li, J.; Yan, Y.; Zhao, X.; Yan, J.; Zhang, Y.; Lai, B.; Chen, X.; Song, L. Catalytic Degradation of Sulfamethoxazole through Peroxymonosulfate Activated with Expanded Graphite Loaded CoFe2O4 Particles. Chem. Eng. J. 2019, 369, 403–413. [Google Scholar] [CrossRef]
  126. Chen, N.; Fang, G.; Zhu, C.; Wu, S.; Liu, G.; Dionysiou, D.D.; Wang, X.; Gao, J.; Zhou, D. Surface-Bound Radical Control Rapid Organic Contaminant Degradation through Peroxymonosulfate Activation by Reduced Fe-Bearing Smectite Clays. J. Hazard. Mater. 2020, 389, 121819. [Google Scholar] [CrossRef] [PubMed]
  127. Tong, Y.; Wang, X.; Zhang, Y.; Xu, J.; Sun, C. Reactive Species in Peracetic Acid-Based AOPs: A Critical Review of Their Formation Mechanisms, Identification Methods and Oxidation Performances. Water Res. 2024, 122917. [Google Scholar] [CrossRef]
  128. 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 2016, 194, 169–201. [Google Scholar] [CrossRef]
  129. Zhang, L.; Jiang, X.; Zhong, Z.; Tian, L.; Sun, Q.; Cui, Y.; Lu, X.; Zou, J.; Luo, S. Carbon Nitride Supported High-loading Fe Single-atom Catalyst for Activation of Peroxymonosulfate to Generate 1O2 with 100% Selectivity. Angew. Chem. Int. Ed. 2021, 60, 21751–21755. [Google Scholar] [CrossRef]
  130. Yu, J.; Tang, L.; Pang, Y.; Liang, X.; Lu, Y.; Feng, H.; Wang, J.; Deng, L.; Zou, J.; Zhu, X. Non-Radical Oxidation in Environmental Catalysis: Recognition, Identification, and Perspectives. Chem. Eng. J. 2022, 433, 134385. [Google Scholar] [CrossRef]
  131. Meng, L.; Dong, J.; Chen, J.; Li, L.; Huang, Q.; Lu, J. Activation of Peracetic Acid by Spinel FeCo2O4 Nanoparticles for the Degradation of Sulfamethoxazole. Chem. Eng. J. 2023, 456, 141084. [Google Scholar] [CrossRef]
  132. 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]
  133. Zeng, Y.; Fang, G.; Fu, Q.; Peng, F.; Wang, X.; Dionysiou, D.D.; Guo, J.; Gao, J.; Zhou, D.; Wang, Y. Mechanistic Study of the Effects of Agricultural Amendments on Photochemical Processes in Paddy Water during Rice Growth. Environ. Sci. Technol. 2022, 56, 4221–4230. [Google Scholar] [CrossRef]
  134. Wu, Z.; Xiong, Z.; Liu, R.; He, C.; Liu, Y.; Pan, Z.; Yao, G.; Lai, B. Pivotal Roles of N-Doped Carbon Shell and Hollow Structure in Nanoreactor with Spatial Confined Co Species in Peroxymonosulfate Activation: Obstructing Metal Leaching and Enhancing Catalytic Stability. J. Hazard. Mater. 2022, 427, 128204. [Google Scholar] [CrossRef] [PubMed]
  135. 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]
  136. Xie, Z.-H.; He, C.-S.; Pei, D.-N.; Dong, Y.; Yang, S.-R.; Xiong, Z.; Zhou, P.; Pan, Z.-C.; Yao, G.; Lai, B. Review of Characteristics, Generation Pathways and Detection Methods of Singlet Oxygen Generated in Advanced Oxidation Processes (AOPs). Chem. Eng. J. 2023, 468, 143778. [Google Scholar] [CrossRef]
  137. Huang, D.; Ren, Y.; Ren, M.; Wang, G.; Du, L.; Li, R.; Xu, W.; Huang, H.; Li, S.; Shen, L. Insights into the Relationship between Regulation of Oxygen Vacancy and Singlet Oxygen Generation in Peracetic Acid Activation. Sep. Purif. Technol. 2025, 359, 130810. [Google Scholar] [CrossRef]
  138. Li, W.; Tang, R.; Xiong, S.; Li, L.; Zhou, Z.; Su, L.; Gong, D.; Deng, Y. High-Valent Metal-Oxo Species in Catalytic Oxidations for Environmental Remediation and Energy Conversion. Coord. Chem. Rev. 2024, 510, 215840. [Google Scholar] [CrossRef]
  139. Zhou, L.; Sleiman, M.; Ferronato, C.; Chovelon, J.-M.; Richard, C. Reactivity of Sulfate Radicals with Natural Organic Matters. Environ. Chem. Lett. 2017, 15, 733–737. [Google Scholar] [CrossRef]
  140. 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]
  141. Yun, E.-T.; Lee, J.H.; Kim, J.; Park, H.-D.; Lee, J. Identifying the Nonradical Mechanism in the Peroxymonosulfate Activation Process: Singlet Oxygenation versus Mediated Electron Transfer. Environ. Sci. Technol. 2018, 52, 7032–7042. [Google Scholar] [CrossRef] [PubMed]
  142. Koppenol, W.H.; Liebman, J.F. The Oxidizing Nature of the Hydroxyl Radical. A Comparison with the Ferryl Ion (FeO2+). J. Phys. Chem. 1984, 88, 99–101. [Google Scholar] [CrossRef]
  143. Yang, B.; Liu, H.; Zhang, J. High-Valent Metals in Advanced Oxidation Processes: A Critical Review of Their Identification Methods, Formation Mechanisms, and Reactivity Performance. Chem. Eng. J. 2023, 460, 141796. [Google Scholar] [CrossRef]
  144. Popova, T.V.; Aksenova, N.V. Complexes of Copper in Unstable Oxidation States. Russ. J. Coord. Chem. 2003, 29, 743–765. [Google Scholar] [CrossRef]
  145. Kang, H.; Lee, D.; Lee, K.-M.; Kim, H.-H.; Lee, H.; Kim, M.S.; Lee, C. Nonradical Activation of Peroxymonosulfate by Hematite for Oxidation of Organic Compounds: A Novel Mechanism Involving High-Valent Iron Species. Chem. Eng. J. 2021, 426, 130743. [Google Scholar] [CrossRef]
  146. Zhang, J.; Xiang, H.; Li, S.; Wang, D.; Yang, B. A Review on the Role of High-Valent Metals in Peracetic Acid-Based Advanced Oxidation Processes. Desalination Water Treat. 2024, 100182. [Google Scholar] [CrossRef]
  147. Wang, Z.; Qiu, W.; Pang, S.; Guo, Q.; Guan, C.; Jiang, J. Aqueous Iron (IV)–Oxo Complex: An Emerging Powerful Reactive Oxidant Formed by Iron (II)-Based Advanced Oxidation Processes for Oxidative Water Treatment. Environ. Sci. Technol. 2022, 56, 1492–1509. [Google Scholar] [CrossRef] [PubMed]
  148. Li, X.; Zhu, W.; Sun, S.-P. Peracetic Acid-Based UVA Photo-Fenton Reaction: Dominant Role of High-Valent Iron Species toward Efficient Selective Degradation of Emerging Micropollutants. J. Hazard. Mater. 2023, 454, 131448. [Google Scholar] [CrossRef]
  149. Zong, Y.; Guan, X.; Xu, J.; Feng, Y.; Mao, Y.; Xu, L.; Chu, H.; Wu, D. Unraveling the Overlooked Involvement of High-Valent Cobalt-Oxo Species Generated from the Cobalt (II)-Activated Peroxymonosulfate Process. Environ. Sci. Technol. 2020, 54, 16231–16239. [Google Scholar] [CrossRef] [PubMed]
  150. Pestovsky, O.; Bakac, A. Reactivity of Aqueous Fe (IV) in Hydride and Hydrogen Atom Transfer Reactions. J. Am. Chem. Soc. 2004, 126, 13757–13764. [Google Scholar] [CrossRef]
  151. Jacobsen, F.; Holcman, J.; Sehested, K. Reactions of the Ferryl Ion with Some Compounds Found in Cloud Water. Int. J. Chem. Kinet. 1998, 30, 215–221. [Google Scholar] [CrossRef]
  152. Wang, Z.; Jiang, J.; Pang, S.; Zhou, Y.; Guan, C.; Gao, Y.; Li, J.; Yang, Y.; Qiu, W.; Jiang, C. Is Sulfate Radical Really Generated from Peroxydisulfate Activated by Iron (II) for Environmental Decontamination? Environ. Sci. Technol. 2018, 52, 11276–11284. [Google Scholar] [CrossRef]
  153. Ao, X.; Wang, W.; Sun, W.; Lu, Z.; Li, C. Degradation and Transformation of Norfloxacin in Medium-Pressure Ultraviolet/Peracetic Acid Process: An Investigation of the Role of PH. Water Res. 2021, 203, 117458. [Google Scholar] [CrossRef]
  154. Kim, K.-H.; Jahan, S.A.; Lee, J.-T. Exposure to Formaldehyde and Its Potential Human Health Hazards. J. Environ. Sci. Health Part. C 2011, 29, 277–299. [Google Scholar] [CrossRef] [PubMed]
  155. Zhang, T.; Wang, T.; Mejia-Tickner, B.; Kissel, J.; Xie, X.; Huang, C.-H. Inactivation of Bacteria by Peracetic Acid Combined with Ultraviolet Irradiation: Mechanism and Optimization. Environ. Sci. Technol. 2020, 54, 9652–9661. [Google Scholar] [CrossRef] [PubMed]
  156. Gasim, M.F.; Bao, Y.; Elgarahy, A.M.; Osman, A.I.; Ala’’a, H.; Rooney, D.W.; Yap, P.-S.; Oh, W.-D. Peracetic Acid Activation Using Heterogeneous Catalysts for Environmental Decontamination: A Review. Catal. Commun. 2023, 180, 106702. [Google Scholar] [CrossRef]
  157. Yao, G.; Xiao, S.; Qian, Y.; Liu, T.; Shi, Y.; Zhang, Y.; Chen, J.; Zhou, X. Non-Metallic Lewis Acid Sites Enhance the Activation of Peracetic Acid by Cobalt-Iron Spinel for Micropollutant Decontamination: The Critical Role of Lewis Acid Sites as Electron Shuttles. Chem. Eng. J. 2024, 496, 153924. [Google Scholar] [CrossRef]
  158. Zheng, L.; Fu, J.; Hua, B.; Wu, Y.; Gu, Y.; Qin, N.; Li, F. Hierarchical Porous Bimetallic FeMn Metal–Organic Framework Gel for Efficient Activation of Peracetic Acid in Antibiotic Degradation. ACS Environ. Au 2023, 4, 56–68. [Google Scholar] [CrossRef]
Figure 1. Potential of PAA activation for achieving simultaneous pathogen inactivation and CECs degradation [19].
Figure 1. Potential of PAA activation for achieving simultaneous pathogen inactivation and CECs degradation [19].
Catalysts 15 00469 g001
Figure 2. Principal reactions involved in SMX removal within the Co/PAA system [77].
Figure 2. Principal reactions involved in SMX removal within the Co/PAA system [77].
Catalysts 15 00469 g002
Figure 3. Potential RS formed in PAA-based AOPs.
Figure 3. Potential RS formed in PAA-based AOPs.
Catalysts 15 00469 g003
Figure 4. Illustration of degradation spectra upon the addition of TBA to identify the presence of H O [21].
Figure 4. Illustration of degradation spectra upon the addition of TBA to identify the presence of H O [21].
Catalysts 15 00469 g004
Table 1. A summary of studies utilizing transition metals for PAA activation in the removal of CECs.
Table 1. A summary of studies utilizing transition metals for PAA activation in the removal of CECs.
CECProcesses[CEC]0Removal Efficiency (%)PAA DoseCatalyst DosagepHReaction TimeRef.
P-arsanilic acidFe2+/PAA5 µM98%400 µM200 µM320 s[63]
Bisphenol-AFe2+/PAA60 mg·L−1100526 µM400 µM3.510 min[64]
NaproxenFe2+/PAA15 µM100100 μM100 μM3–8.2120 min[39]
Rhodamine BCoFe2O4/PAA20 mg·L−195800 μM2131.1 μM710 min[65]
SulfamethoxazoleFe2+-zeolite/PAA5 µM100400 μM150 mg·L−1750 min[62]
TetracyclineFe0/PAA10 μM95100 μM1074.4 µM3.530 min[66]
SulfamethoxazoleCo2+/PAA10 μM100100 μM10 μM3.520 min[41]
CarbamazepineCo2+/PAA10 μM84100 μM10 μM3.530 min[41]
Bisphenol-ACo2+/PAA10 μM100100 μM10 μM3.515 min[41]
AtrazineCo2+/PAA10 μM20100 μM10 μM3.530 min[41]
NaproxenCo2+/PAA15 µM100100 μM10 μM3–8.13 min[20]
Crystal VioletCo2+/PAA0.06 mM67200 μM10 μM760 min[67]
Congo redCo2+/PAA0.05 mM98200 μM10 μM760 min[67]
Acid orange 7Co2+/PAA0.05 mM92200 μM10 μM760 min[67]
Orange GCo3O4/PAA0.05 mM1006574.6 μM415.3 μM790 min [68]
SulfamethoxazoleCoFe2O4/PAA10 μM85100–200 μM426.2 μM730 min[52]
SulfamethoxazoleLaCoO3/PAA50 μM100263 μM660 μM760 min[69]
SulfamethoxazoleCo0/PAA5 μM9950 μM1696.8 μM75 min[30]
SulfamethoxazoleMn3O4/PAA1 μM1001000 μM218.5 μM6.512 min[70]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Bouzayani, B.; Elaoud, S.C.; Sanromán, M.Á. Current Progress in Advanced Oxidation Processes for the Removal of Contaminants of Emerging Concern Using Peracetic Acid as an Effective Oxidant. Catalysts 2025, 15, 469. https://doi.org/10.3390/catal15050469

AMA Style

Bouzayani B, Elaoud SC, Sanromán MÁ. Current Progress in Advanced Oxidation Processes for the Removal of Contaminants of Emerging Concern Using Peracetic Acid as an Effective Oxidant. Catalysts. 2025; 15(5):469. https://doi.org/10.3390/catal15050469

Chicago/Turabian Style

Bouzayani, Bakhta, Sourour Chaâbane Elaoud, and Maria Ángeles Sanromán. 2025. "Current Progress in Advanced Oxidation Processes for the Removal of Contaminants of Emerging Concern Using Peracetic Acid as an Effective Oxidant" Catalysts 15, no. 5: 469. https://doi.org/10.3390/catal15050469

APA Style

Bouzayani, B., Elaoud, S. C., & Sanromán, M. Á. (2025). Current Progress in Advanced Oxidation Processes for the Removal of Contaminants of Emerging Concern Using Peracetic Acid as an Effective Oxidant. Catalysts, 15(5), 469. https://doi.org/10.3390/catal15050469

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