You are currently viewing a new version of our website. To view the old version click .
Nanomaterials
  • Review
  • Open Access

12 November 2025

Iron Redox Cycling in Persulfate Activation: Strategic Enhancements, Mechanistic Insights, and Environmental Applications—A Review

,
,
,
and
1
College of Environmental Science and Engineering, Tongji University, Shanghai 200092, China
2
Circular Economy Research Institute, School of Marxism, Tongji University, Shanghai 200092, China
3
Yangtze River Delta Institute of Circular Economy Technology, Jiaxing 314001, China
4
State Key Laboratory of Pollution Control and Resources Reuse, College of Environmental Science and Engineering, Tongji University, Shanghai 200092, China
Nanomaterials2025, 15(22), 1712;https://doi.org/10.3390/nano15221712 
(registering DOI)
This article belongs to the Special Issue Advances in Nanomaterials for Energy Conversion and Environmental Catalysis: Second Edition

Abstract

Iron-based catalysts for peroxymonosulfate (PMS) and peroxydisulfate (PDS) activation represent a cornerstone of advanced oxidation processes (AOPs) in environmental remediation, prized for their cost-effectiveness, environmental compatibility, and high catalytic potential. These catalysts, including zero-valent iron, iron oxides, and iron-organic frameworks, activate PMS/PDS through heterogeneous and homogeneous pathways to generate reactive species such as sulfate radicals (SO4) and hydroxyl radicals (•OH). However, their large-scale implementation is constrained by inefficient iron cycling, characterized by sluggish Fe3+/Fe2+ conversion and significant iron precipitation, leading to catalyst passivation and oxidant wastage. This comprehensive review systematically dissects innovative strategies to augment iron cycling efficiency, encompassing advanced material design through elemental doping, heterostructure construction, and defect engineering; system optimization via reductant incorporation, bimetallic synergy, and pH modulation; and external field assistance using light, electricity, or ultrasound. We present a mechanistic deep-dive into these approaches, emphasizing facilitated electron transfer, suppression of iron precipitation, and precise regulation of radical versus non-radical pathways. The performance in degrading persistent organic pollutants—including antibiotics, per- and polyfluoroalkyl substances (PFASs), and pesticides—in complex environmental matrices is critically evaluated. We further discuss practical challenges related to scalability, long-term stability, and secondary environmental risks. Finally, forward-looking directions are proposed, focusing on rational catalyst design, integration of sustainable processes, and scalable implementation, thereby providing a foundational framework for developing next-generation iron-persulfate catalytic systems.

1. Introduction

The rapid advancement of industrialization and urbanization has led to the widespread contamination of water and soil environments by refractory organic pollutants, including antibiotics, dyes, pesticides, and PFAS [,,]. These contaminants pose severe risks to ecosystems and human health due to their high toxicity, chemical stability, and low biodegradability. Conventional treatment techniques—such as biodegradation, adsorption, and physical separation—often prove ineffective for complete removal and may entail high operational costs or secondary pollution concerns. Consequently, there is an urgent need to develop efficient, economical, and environmentally sustainable degradation technologies.
Advanced oxidation processes (AOPs) have emerged as a promising strategy for deep mineralization of organic pollutants through the generation of reactive oxygen species (ROS). However, conventional Fenton reactions suffer from inherent limitations, including a narrow operational pH range (typically 2–4), substantial iron sludge production, and risks associated with the storage and transportation of H2O2 [,,]. These drawbacks have motivated the development of novel AOPs that exhibit high efficiency, broad pH applicability, and minimal environmental impact.
Persulfate-based AOPs, utilizing peroxymonosulfate (PMS) or peroxydisulfate (PDS), have garnered considerable interest as viable alternatives. Persulfates offer advantages such as high stability, facilitating safer handling and transport, as well as high redox potentials that enable effective degradation of various recalcitrant organics [,]. Upon activation, persulfates generate multiple reactive species, primarily SO4 and •OH, which exhibit strong oxidative capacity and broad substrate applicability. Despite their potential, persulfates alone are incapable of degrading pollutants; they require catalytic activation to cleave the peroxy bond and produce ROS. Activation methods include transition metal catalysis, carbon-based materials, and energy-based processes such as UV and heating [,,,].
Among these, iron-based catalysts—such as zero-valent iron (ZVI), iron oxides, and iron-organic frameworks—are regarded as ideal activators due to the abundance, low cost, and low toxicity of iron [,,,,]. Nevertheless, their catalytic efficiency is severely hampered by low iron cycling efficiency [,,,,]. While Fe2+ rapidly activates PMS/PDS to yield SO4•−/•OH, it is subsequently oxidized to Fe3+ (with rate constants k in the range of 10–100 M−1s−1). The reduction of Fe3+ back to Fe2+ is notably slow (k < 0.001–1 M−1s−1), particularly in heterogeneous systems [,,,,,]. Moreover, under neutral or alkaline conditions, Fe3+ tends to form iron (oxyhydr)oxide precipitates (e.g., FeOOH, Fe(OH)3), which encapsulate active sites and impede the catalytic cycle. These issues result in inefficient oxidant utilization, incomplete pollutant mineralization, and limited pH applicability (often requiring acidic conditions), significantly hindering practical implementation.
There has been a surge of research interest in persulfate activation using iron-based catalysts. Bibliometric analysis of the Web of Science Core Collection reveals a marked upward trend in annual publications focused on iron-persulfate systems, underscoring the growing importance of this technology (Figure 1a). Keyword co-occurrence network analysis further indicates that enhancing iron cycling efficiency is a core research direction in iron-mediated persulfate-based AOPs to achieve the improvement of process performance and efficient removal of organic pollutants (Figure 1b). This growing research interest has yielded significant advances in understanding and enhancing the iron redox cycle. Recent studies have demonstrated that atomic-level engineering through single-atom Fe-N4 configurations can achieve exceptional persulfate activation while minimizing metal leaching []. Concurrently, heterointerface construction, such as in MoS2/Fe3O4 composites, has proven effective in facilitating interfacial electron transfer and accelerating Fe3+/Fe2+ conversion []. Defect engineering through the creation of oxygen or sulfur vacancies has emerged as another powerful strategy to modulate local electron density and optimize persulfate adsorption [,]. Furthermore, sustainable approaches utilizing waste-derived materials, such as catalysts synthesized from spent lithium-ion batteries, have shown promising performance while addressing environmental sustainability concerns [,]. These developments represent the cutting edge of iron redox cycling research and provide deeper mechanistic insights into the interfacial processes governing persulfate activation.
Figure 1. (a) Annual number of publications on iron-catalyzed persulfate activation for environmental remediation, indicating consistent growth and increasing research interest since 2015. (b) Keyword co-occurrence network analysis visualizing major research themes and conceptual relationships in iron-enhanced persulfate activation systems.
Despite the existence of previous reviews on iron-based persulfate activation, this work distinguishes itself by providing a unified and critical analysis centered explicitly on overcoming the iron cycling bottleneck. Unlike prior works that often catalog materials or applications, this review establishes a comprehensive “problem identification-strategy development-mechanism elucidation-application evaluation” framework. Our unique synthesis integrates enhancement strategies across catalyst design, system optimization, and external field assistance into a single mechanistic narrative, revealing how disparate approaches ultimately converge on accelerating electron transfer and suppressing iron deactivation. Furthermore, we place a distinct emphasis on establishing structure-activity relationships and on the practical implications of these mechanisms in complex environmental matrices, topics often treated superficially. This review aims to provide not only a summary of the state-of-the-art but also a foundational guide for the rational design of next-generation iron-persulfate systems.

2. The Root Causes and Impacts of Low Iron Cycling Efficiency

2.1. Thermodynamic and Kinetic Limitations

The inefficiency of the Fe2+/Fe3+ redox cycle, central to persulfate activation, arises from deeply intertwined thermodynamic and kinetic constraints (Figure 2). Thermodynamically, the reduction of Fe3+ to Fe2+ must overcome a substantial energy barrier, as reflected in its high standard reduction potential (E°(Fe3+/Fe2+) = +0.77 V vs. NHE). In persulfate-driven systems, the frequent absence of efficient endogenous electron donors promotes the hydrolysis of Fe3+ into insoluble and thermodynamically stable precipitates, such as Fe(OH)3 or FeOOH, which exhibit exceedingly low solubility constants (Ksp ≈ 10−39). This precipitation effectively depletes the dissolved Fe2+ pool, trapping the iron cycle in a kinetically inert, oxidized state. Furthermore, although high-valent iron species (e.g., Fe(IV)=O) possess high oxidative potential, their generation requires overcoming considerable activation energy barriers and often proves unsustainable without appropriate ligand stabilization or confined coordination environments [,,]. Essentially, the system becomes thermodynamically “locked” in the Fe3+ state, stalling persistently in the absence of an efficient electron-compensation mechanism.
Figure 2. The Root Causes and Impacts of Low Iron Cycling Efficiency.
Kinetically, the iron cycle is impeded by three major bottlenecks. First, in homogeneous systems, the apparent rate constant (k) for the reduction of Fe3+ to Fe2+ decreases by 2–3 orders of magnitude as pH increases beyond 5, largely due to the predominance of less reactive hydrolyzed species such as Fe(OH)2+ []. Second, in heterogeneous catalytic systems, the accumulation of passivating layers (e.g., FeOOH) leads to an exponential increase in charge transfer resistance (often exceeding 10-fold after multiple cycles), severely impeding electron transport from the bulk material to reactive surface sites [,]. Third, radical scavenging introduces a self-consuming pathway: sulfate (SO4•−) and hydroxyl (•OH) radicals react with Fe2+ at near diffusion-limited rates (k ≈ 109 M−1 s−1), significantly outpacing their reactions with target pollutants []. This leads to rapid consumption of Fe2+ on a microsecond timescale, macroscopically manifesting as a “kinetic short-circuit” within the iron cycle. Together, these kinetic barriers create a “rate trap” that drastically diminishes catalytic efficiency, even under otherwise thermodynamically favorable conditions.
The inhibition of iron cycling triggers a rapid decline in the system oxidation-reduction potential (ORP), yielding several critical consequences: (1) A drastic reduction in oxidant utilization: Depletion of Fe2+ causes persulfate activation rates to drop by 80–90%, resulting in an exponential decay in SO4 yield and extending the half-lives of pollutant degradation by 1–2 orders of magnitude [,,]. (2) A shift in reaction pathways: The proportion of high-valent iron species (e.g., Fe(IV)=O) and non-radical oxidation pathways increases. Although capable of degrading certain contaminants, these pathways often exhibit poor selectivity toward chlorinated and nitrogenous organics, frequently yielding more toxic intermediates (e.g., chlorinated quinones and nitro-byproducts) [,,]. (3) A cascade of ecotoxicological effects: Fe(OH)3 flocs can adsorb toxic transformation products, forming “iron-toxin” co-precipitates that accumulate in sediments and pose a secondary pollution risk. Moreover, these flocs may catalyze the formation of persistent free radicals (PFRs), initiating secondary chain reactions that further amplify environmental toxicity [,,,]. In summary, inefficient iron cycling not only compromises treatment performance but also exacerbates ecological risks, creating a detrimental negative feedback loop between reduced efficiency and heightened toxicity.

2.2. Challenges of Fe3+ Precipitation

Iron precipitation is a major contributor to the inefficient iron cycling observed in iron-based advanced oxidation processes. Common precipitates include iron (oxyhydr)oxides (e.g., FeOOH, Fe(OH)3), mixed-valence oxides (e.g., Fe3O4, γ-Fe2O3), and co-precipitates with anions such as phosphates or carbonates. These precipitates impede the iron cycle through multiple concurrent mechanisms. Primarily, precipitate formation drastically reduces the accessibility of active iron sites. For instance, a goethite (α-FeOOH) shell thicker than 3 nm can completely block Fe2+ diffusion, increasing charge transfer resistance (Rₜ) by 1–2 orders of magnitude and severely diminishing redox cycling efficiency [,]. Secondly, precipitation alters local iron redox thermodynamics. The formation of FePO4, for example, depresses the standard Fe3+/Fe2+ redox potential from +0.77 V to +0.36 V, thermodynamically suppressing Fe2+ regeneration []. Additionally, surface-adsorbed Fe(II) species are rapidly oxidized by sulfate radicals (SO4•−; k ≈ 4.6 × 109 M−1s−1), disrupting the balance between surface and solution iron cycles and macroscopically reducing oxidant utilization by 70–80% [,].
The formation of iron precipitates is influenced by several environmental and operational factors, including pH, anion concentration, and catalytic cycling number. For example, as pH increases from 5 to 8, the formation rate of FeOOH rises significantly, with a pseudo-first-order precipitation rate constant (k) reaching 0.03 min−1 []. High concentrations of H2PO4 and HCO3 promote the formation of FePO4 and FeCO3 co-precipitates, markedly decreasing Fe(II) availability and further hindering the catalytic cycle []. With repeated catalytic cycles, the accumulation of precipitates on the catalyst surface leads to progressive deactivation, posing a significant challenge for long-term operation.
In summary, iron precipitation severely impedes the iron cycle through two interconnected mechanisms: (1) “thermodynamic locking”, where the formation of stable, insoluble iron (oxyhydr)oxide precipitates (e.g., FeOOH, Fe(OH)3) shifts the system equilibrium, making the reduction of Fe3+ back to Fe2+ thermodynamically less favorable [,]; and (2) “kinetic shielding”, where these precipitates form a physical barrier on the catalyst surface, blocking active sites and drastically increasing the resistance to electron transfer, thereby kinetically hindering the Fe3+/Fe2+ conversion [,]. To counter these issues, strategies such as ligand complexation, surface defect engineering, and interface modulation have proven effective in mitigating precipitation and restoring rapid iron cycling (Table 1). These approaches collectively enhance both the efficiency of the iron cycle and the utilization of oxidants, providing crucial theoretical insights and practical strategies for deploying iron-based advanced oxidation systems in complex water environments. The following sections will elaborate on these enhancement strategies in detail.
Table 1. Fundamental Limitations Governing Inefficient Iron Cycling in Persulfate Activation Systems.

3. Iron Cycle Enhancement Strategies and Their Mechanisms

3.1. Catalyst Design and Modification Strategies for Enhanced Iron Cycling

The strategic design of high-performance iron-based catalysts, as detailed in the following subsections, is intrinsically linked to their synthesis. The implementation of doping, heterointerfaces, defects, and morphological control is achieved through a suite of synthetic methods, each with its own merits and limitations []. Common techniques include (i) hydrothermal/solvothermal routes for high-crystallinity nanomaterials with defined morphologies, (ii) pyrolysis for carbon-based composites and single-atom catalysts, (iii) co-precipitation for scalable production of mixed metal oxides, and (iv) template-assisted methods for precise architectural control. The selection of a specific route (e.g., sol-gel, ball milling) is dictated by the target material’s properties, balancing factors such as scalability, cost, and the need for precise control over crystallinity, porosity, and elemental distribution [,].

3.1.1. Electronic Structure Modulation via Elemental Doping

Elemental doping represents a sophisticated strategy to overcome the kinetic constraints of the Fe3+/Fe2+ cycle by deliberately engineering the electronic and coordination structures of iron-based catalysts (Figure 3). This strategy primarily operates through three distinct yet complementary mechanisms: (i) The incorporation of transition metals (e.g., Co, Cu, Mo) establishes coupled redox cycles where the dopant’s valence cycle (e.g., Co2+/Co3+) acts as an efficient electron-transfer mediator, shuttling electrons to Fe3+ and bypassing its slow direct reduction. (ii) Non-metal dopants (e.g., S, N, B) primarily modify the electronic structure of the support matrix (e.g., carbon), enhancing electron density and creating defined coordination environments to stabilize iron active sites and lower activation energies. (iii) Dual-element doping schemes synergize these effects, combining rapid metal redox cycling with an optimized electron-conducting matrix to further enhance performance and pathway selectivity. This approach introduces foreign elements—including both transition metals and non-metals—that actively participate in or facilitate electron transfer processes, thereby accelerating persulfate activation and sustaining catalytic longevity (Figure 4a) (Table 2).
Figure 3. Electronic structure modulation via elemental doping to produce ROS.
Figure 4. Catalyst Design and Modification Strategies for Enhanced Iron Cycling. (a) Elemental Doping. (b) Heterointerface Engineering. (c) Defect Engineering. (d) Morphological Engineering.
The incorporation of transition metals (e.g., Co, Cu, Mo) establishes coupled redox systems wherein the dopant’s valence cycle (e.g., Co2+/Co3+, Cu+/Cu2+) acts as an efficient electron-transfer mediator. For instance, in Co-Fe3O4@FeOOH, the Co2+/Co3+ couple facilitates interfacial electron donation to Fe3+, enhancing the methylene blue degradation rate by 36-fold while restricting cobalt leaching to below 0.457 mg/L []. Beyond electron shuttling, certain dopants modulate the electronic properties of iron active sites. Density functional theory (DFT) calculations reveal that Mo doping in Fe/Mo@C downshifts the d-band center of iron, strengthening PMS adsorption and reducing the energy barrier for O–O bond cleavage which culminates in a 26-fold increase in bisphenol A degradation rate [].
Non-metal dopants (e.g., S, N, B) primarily modify the support matrix to enhance electron density and create defined coordination environments (Figure 3). Electron-rich nitrogen configurations (e.g., pyrrolic N) in Fe-N-BC donate electron density to iron centers, stabilizing Fe2+ and lowering the activation energy for persulfate dissociation (ΔG = 23.54 kcal/mol), achieving 90.2% sulfamethoxazole degradation within 40 min []. Sulfur doping introduces reversible redox-active sites (e.g., S2−/S0) that directly reduce Fe3+, as demonstrated in Fe/S-NC, where thiophenic sulfur promotes interfacial electron polarization and sustains Fe2+ regeneration, maintaining 96.3% activity after five cycles [].
Dual-element doping schemes yield synergistic effects by combining rapid metal redox cycling with enhanced electron transport through a modified carbon matrix. In FeCo@NBC, cobalt accelerates iron reduction while nitrogen doping increases graphitization (Iₚ/I_G = 1.04), collectively promoting a singlet oxygen-dominated pathway and high degradation efficiency for bisphenol S []. Crucially, dopant identity dictates reaction mechanism: sulfur incorporation in Fe@S4NC shifts the dominant reactive species from singlet oxygen to sulfate radicals, elevating sulfadiazine mineralization by 32% [].
In essence, elemental doping enhances iron cycling through three convergent mechanisms: coupled redox mediation [], electronic structure tailoring [], and created defined coordination environments [,]. This strategy provides a foundational methodology for circumventing kinetic bottlenecks in iron-persulfate catalysis through targeted materials design. However, the choice of dopant is not arbitrary but should be guided by the desired electronic modulation and catalytic function. General principles for rational dopant selection are emerging from combined experimental and theoretical studies: (1) For enhancing electron transfer, transition metals with compatible and rapid redox cycles (e.g., Co2+/Co3+, Cu+/Cu2+) relative to Fe3+/Fe2+ are preferred. Their efficacy can be predicted by comparing standard reduction potentials and the energy barriers for electron transfer calculated via Density Functional Theory (DFT). (2) For modulating the electronic structure of the iron center or carbon support, non-metals with distinct electronegativities (e.g., N, S, B) are chosen to create electron-deficient or electron-rich regions, which can be tailored to optimize persulfate adsorption and activation. (3) The selection of dopant precursors is critical. Metal dopants often originate from nitrate, chloride, or acetylacetonate salts, chosen for their solubility and decomposition behavior during synthesis. Non-metal dopants are typically introduced from nitrogen-rich (e.g., urea, melamine), sulfur-rich (e.g., thiourea, L-cysteine), or boron-containing (e.g., boric acid) precursors. The selection is based on the precursor’s ability to incorporate efficiently into the host matrix at the synthesis temperature and its impact on the final material’s morphology and defect density. Ultimately, the integration of DFT calculations—predicting parameters like d-band center, adsorption energy, and charge density distribution—prior to experimentation is becoming an indispensable tool for the rational design of doped catalysts.

3.1.2. Heterointerface Engineering for Enhanced Electron Transfer

Constructing heterostructures represents a sophisticated materials design strategy to overcome electron-transfer limitations in iron-based persulfate activation (Figure 4b). By deliberately engineering interfaces between iron compounds and carefully selected secondary phases—such as conductive carbons or other transition metal oxides—this approach creates synergistic effects that markedly accelerate the Fe3+/Fe2+ cycle and mitigate deactivation processes [].
The integration of carbonaceous materials establishes efficient electron-transfer pathways that facilitate iron valence cycling. In core–shell Fe3O4@NCNTs, the graphitic carbon layers serve as conductive networks, tripling the rate of Fe3+ reduction while significantly suppressing iron leaching []. Confinement of Fe2O3 within oxidized carbon nanotubes (Fe2O3@OCNT) enhances electronic interaction and increases PMS adsorption, achieving 96.1% tetracycline degradation with minimal iron leaching (<10 μg/L) []. Biochar-based systems leverage inherent functional groups (e.g., C=O, pyridinic N) and defect sites to improve iron dispersion and electron shuttling. The synergistic effect between graphitic carbon and Fe–Nₓ sites in FeN@BC enables complete tetracycline degradation within 30 min across a broad pH range (3–9) [].
Introducing a second metal species creates bimetallic redox pairs that establish efficient electron-relay mechanisms. In a Fe–Cu dual-atom catalyst, Cu+ serves as an electron shuttle—preferentially reducing Fe3+ while oxidizing persulfate—resulting in a 40% higher Fe2+ regeneration rate and a sixfold increase in PDS activation []. In Fe–Mn@biochar systems, Mn3+/Mn4+ cycles drive Fe2+ regeneration and contribute to 70% of radical generation, boosting the degradation rate constant by 20 times compared to monometallic catalysts [].
Beyond enhancing electron transfer, heterostructures exert precise control over reaction pathway selection. Confinement effects and tailored interfaces can favor non-radical routes that are less susceptible to quenching by complex water matrices. For instance, Fe–N4 sites in single-atom catalysts activate PMS via direct electron transfer, while oxygen vacancy-rich ZnFe2O4/CNTs promote 1O2 generation with 89% contribution [,]. Encapsulation strategies, such as Fe3C encapsulated in graphitic carbon (Fe3C@C), reduce metal leaching by 90%, effectively decoupling high activity from stability losses [].
In summary, heterointerface engineering enhances persulfate activation through three complementary mechanisms: the electronic pole, where conductive interfaces lower energy barriers for Fe3+ reduction; the synergistic pole, where bimetallic relays enable continuous electron shuttling; and the pathway pole, involving confinement-induced control over radical and non-radical pathways. This multi-level strategy provides a universal blueprint for designing robust iron-based catalysts with enhanced activity and stability. The rational design of such interfaces is paramount. The selection of secondary phases for constructing heterointerfaces follows a similar rationale driven by desired functionality. Conductive carbons (e.g., graphene, carbon nanotubes) are chosen to provide high-speed electron pathways, while secondary metal oxides or sulfides are selected to establish synergistic redox cycles or to create built-in electric fields that enhance charge separation. The choice is increasingly guided by theoretical predictions of interfacial energy, band alignment, and electron affinity at the junction, ensuring the designed interface facilitates the directional electron transfer crucial for Fe3+ reduction.

3.1.3. Defect Engineering for Regulating Electron Transfer and Reaction Pathways

Defect engineering has emerged as a pivotal strategy for modulating the electronic structure of iron-based catalysts to overcome kinetic limitations in persulfate activation. By deliberately introducing vacancies—such as oxygen vacancies (Ov) and sulfur vacancies (Sv)—this approach optimizes the local coordination environment and electronic properties of metal active sites, thereby enhancing electron transfer efficiency, regulating reaction pathways, and improving catalytic stability (Figure 4c).
Oxygen vacancies function as electron reservoirs that significantly increase electron density at the catalyst surface. In CoWO4−x, Ov-induced localized electron rearrangement promotes direct electron donation to adsorbed PMS molecules, reducing the activation energy for O–O bond cleavage []. Similarly, graphene-supported α-Fe2O3−x exhibits improved electrical conductivity due to Ov, which enhances Fe2+ regeneration and sustains persulfate activation []. Sulfur vacancies play a complementary role by exposing under-coordinated metal sites and optimizing adsorbate-catalyst interactions. In FeS/MoS2 systems, Sv exposes active Mo4+ sites that serve as electron mediators, facilitating Fe3+ reduction and promoting continuous iron cycling [].
Beyond enhancing electron transfer, vacancy defects exert precise control over reaction mechanism orientation. Oxygen vacancies can lower the energy barrier for molecular oxygen adsorption, thereby promoting the generation of singlet oxygen (1O2) []. This is exemplified in MnCeOx–H, where Ov facilitates the formation of metastable PMS surface complexes that decompose preferentially via the 1O2 pathway []. Sulfur vacancies in CoFe2O4−x mediate a stepwise reduction pathway, converting PMS into O2 and subsequently to 1O2 []. Low-coordination defect environments—such as asymmetric oxygen vacancies (As-OVs) in CuZnOx [] and Co-N2 sites in single-atom catalysts—further promote direct electron transfer processes (ETPs) by stabilizing metal–PMS* complexes with optimized orbital overlap [].
Crucially, vacancy engineering addresses the persistent challenge of metal leaching through structural stabilization. Defects serve as structural anchors that stabilize metal centers via lattice confinement and electronic compensation. In CdxZn1−xS@BC, Zn-induced charge redistribution and vacancy formation suppress cadmium oxidation, reducing leaching by 62.8% []. Prolonged catalytic stability is achieved in silicate-confined systems, where iron cycling remains effective over 192 h []. The presence of Ov in FeVO3−x [] and Co3O4@C-500 [] further stabilizes low-valence states (V4+, Co2+), enabling high reactivity over multiple cycles.
In summary, defect engineering enhances persulfate activation through three interconnected mechanisms: electronic facilitation, where vacancies act as electron donors/acceptors to accelerate metal redox cycling; pathway control, where defect-induced electronic restructuring favors specific oxidation routes (1O2 and ETP); and structural stabilization, whereby vacancies anchor metal species, suppress dissolution, and extend catalyst lifespan. This approach provides atomic-level insights for the rational design of highly efficient and stable iron-based catalytic systems.

3.1.4. Morphological and Structural Engineering for Enhanced Mass and Electron Transport

Rational manipulation of catalyst morphology and structure plays a critical role in overcoming intrinsic limitations in persulfate activation by optimizing physical and chemical pathways for mass diffusion and electron transfer (Figure 4d). This approach enhances the accessibility of active sites, facilitates reactant and charge transport, and tailors reaction environments, thereby significantly improving the efficiency and sustainability of iron-based catalytic systems.
Nanostructural engineering increases specific surface area and exposes a greater density of active sites, improving PMS/PS adsorption and activation kinetics. For example, a SiO2-templated and NaNO3-activated porous carbon structure achieved a specific surface area of 773 m2/g. The abundant structural defects and synergistic FexOy-N-C effects enhanced PMS activation, enabling efficient pollutant degradation over a broad pH range (3–9) []. Hollow core–shell architectures, such as Fe3O4@CoFe2O4, utilize an internal cavity to increase reactive surface area while the outer shell protects the ZVI core from passivation, enabling sustained Fe2+ release and extending catalyst lifetime to 22 h [].
Low-dimensional nanostructures—such as two-dimensional nanosheets and one-dimensional nanowires—shorten electron transport distances, thereby accelerating the Fe2+/Fe3+ cycle and improving oxidant utilization efficiency. Ultrathin Fe3O4 nanosheets (≈1 nm thickness) with abundant oxygen vacancies reduced the O–O bond cleavage energy barrier, increasing the PMS activation rate by five times []. MoS2 nanosheets coating Fe3O4 form Fe-S bridging interfaces that accelerate electron transfer and Fe2+ regeneration, achieving a degradation rate four times higher than that of pure Fe3O4 [].
Ordered pore channels and hierarchical structures reduce mass transfer resistance and enhance the diffusion of pollutants and oxidants to active sites. Ordered porous Fe3O4@N–C cubes with acid-etched aligned channels (OA-P structure) accelerated ion/electron transport, simultaneously enhancing the iron cycle and PMS activation efficiency (99.8% BPA degradation within 60 min) []. Hollow flower-like γ-MnO2 loaded with iron was assembled from ultrathin nanosheets; the hierarchical porosity promoted As(III) diffusion to iron active sites, yielding a synergistic oxidation–adsorption capacity of 102.84 mg/g [].
Certain morphological features further optimize catalytic performance through defect engineering or surface reconstruction, steering radical versus non-radical pathways. In MoS2-Fe, sulfur vacancies (Sᵥ) enhanced PMS adsorption and facilitated the formation of a FeIII–PMS* complex, driving non-radical pathways via 1O2 and Fe(IV)=O (contributing 45.5%) []. In core–shell Fe@C-1000, pore formation during Fe0 consumption allowed internal Fe3C to migrate to the surface, replenishing active sites with nearly zero iron leaching (0.08–0.23 mg/L) and shifting the dominant ROS from SO4 to high-valent iron Fe(IV) [].
Morphological and structural regulation enhances persulfate activation through three fundamental mechanisms: improved mass transfer through hierarchical porosity, accelerated electron transport via low-dimensional architectures, and optimized reaction pathways through defect-assisted interface engineering. These strategies collectively address key limitations in iron-based persulfate activation and provide a foundation for designing high-performance catalytic systems.
In summary, the catalyst design strategies discussed in this section collectively address the iron cycling bottleneck through three fundamental mechanisms: electronic structure modulation, interface engineering, and mass/electron transport optimization. Elemental doping introduces foreign elements that create electron-transfer mediators and modify the coordination environment of iron sites, significantly enhancing the Fe3+/Fe2+ conversion kinetics [,]. However, the long-term stability of dopants and potential leaching issues require careful consideration. Heterointerface engineering constructs hybrid interfaces that provide continuous electron-transfer pathways and establish synergistic redox relays between different components [,]. While highly effective, the structural integrity of these interfaces under prolonged operation remains a challenge. Defect engineering creates vacancies that serve as electron reservoirs and modulate surface adsorption/activation properties, enabling precise control over reaction pathways [,]. Nevertheless, the controlled generation and maintenance of these defects under realistic conditions need further investigation. Morphological engineering optimizes physical structures to enhance mass diffusion and active site accessibility, though scalability and practical implementation of complex nanostructures pose significant hurdles [,]. Looking forward, the integration of multiple strategies—such as doped materials with tailored defects and controlled morphology—represents a promising direction for developing robust, high-performance iron-based catalysts that maintain efficiency under environmentally relevant conditions.
Table 2. Comparative Analysis of Catalyst Design Strategies for Enhancing Iron Cycling in Persulfate Activation.
Table 2. Comparative Analysis of Catalyst Design Strategies for Enhancing Iron Cycling in Persulfate Activation.
StrategyPrimary Mechanism & Electronic ModulationKey Structural FeatureDominant Enhancement in Iron CyclingTypical Performance OutcomeReferences
Elemental DopingIntroduces foreign elements to modify the electronic structure and create electron-transfer mediatorsAtomic-level dispersion of dopants; formation of active sites Couples Fe3+/Fe2+ with a faster redox pair; lowers activation energy for Fe3+ reduction.26-fold increase in BPA degradation rate; >96% activity retention after 5 cycles.[,]
Heterointerface Engineering Constructs hybrid interfaces to provide continuous electron-transfer pathways and synergistic redox relays.Core–shell structures; bimetallic junctions; carbon-supported composites.Accelerates interfacial electron transfer from support or second metal to Fe3+.6-fold higher PDS activation; 96% degradation with minimal leaching (<10 μg/L).[,]
Defect Engineering Creates vacancies (Oᵥ, Sᵥ) that act as electron reservoirs and modulate surface adsorption/activation.Oxygen/sulfur vacancies; under-coordinated metal sites; asymmetric coordination.Serves as localized electron donor for Fe3+ reduction; stabilizes low-valence states.62.8% reduction in metal leaching; sustained activity over 192 h.[,]
Morphological EngineeringControls physical structure to shorten mass/electron transport distances and increase active site exposure.Hollow structures; ordered pores; 2D nanosheets; hierarchical porosity.Improves diffusion of reactants to sites and electrons to Fe centers.5-fold higher PMS activation rate; 99.8% degradation in 60 min.[,]

3.2. Reaction System Optimization Strategies

Building upon the intrinsic strategies of catalyst design discussed in Section 3.1—which focused on permanently embedding enhanced properties within the catalyst’s architecture—this section shifts the paradigm to extrinsic system optimization. Here, the enhancement of the iron cycle is achieved not by altering the catalyst’s fundamental identity, but by strategically engineering the external reaction environment. This is accomplished through the introduction of soluble chemical agents or solid supplements that act as auxiliary components within the reaction system. These extrinsic agents work in concert with the catalyst, operating alongside it in the solution or at the interface to dynamically supply electrons, mediate charge transfer, and prevent deactivation. This approach offers a complementary pathway to overcome kinetic limitations, often providing greater operational flexibility and rapid efficacy, albeit sometimes at the cost of long-term reagent consumption or potential secondary effects.

3.2.1. Introduction of Chemical Reductants

The introduction of chemical reductants addresses the intrinsic kinetic limitation of the Fe(III)/Fe(II) cycle by providing exogenous electrons, thereby enabling sustained persulfate activation and dramatically enhancing the degradation efficiency of refractory pollutants (Figure 5a). This strategy operates through direct electron donation, complexation-mediated reduction, or the formation of transient reactive intermediates that facilitate iron valence cycling (Table 3). Unlike the permanent electronic structure modulation achieved by doping, chemical reductants function as soluble, sacrificial electron donors in the solution phase, providing a transient but powerful boost to the iron cycle.
Figure 5. Optimization Strategies via Reductive Enhancement. (a) Chemical Reductant. (b) Electron Shuttles. (c) Elemental Substances.
Inorganic reductants, such as hydroxylamine (HA) and dithionite (DTN), primarily function via direct electron transfer to iron centers. HA enhances atrazine degradation by 38-fold in a Fe3O4/PMS system, serving both as a homogeneous activator and a surface-mediated electron shuttle through Fe–O–N bonding []. DTN exhibits stronger reducibility by generating SO4 radicals, which drive a coupled vanadium-iron redox cycle in FeVO4/PMS systems, achieving >90% carbamazepine removal within 300 s (Lai et al., 2021 []). Sulfite-based reductants (e.g., HSO3) demonstrate remarkable pH adaptability through complexation with dissolved iron, enhancing reactivity via multiple pathways: reduction of Fe(III), reaction with Fe(IV) to produce SO4, and in situ generation of PMS [].
Organic reductants offer the dual function of reducing Fe(III) and solubilizing iron species, thereby preventing precipitation and maintaining iron bioavailability. Ascorbic acid (AA) significantly enhances sulfamethazine degradation (72.4% improvement) in Fe(III)/PDS systems by reducing surface-bound Fe(III) and forming soluble [Si–Fe(II)] complexes, minimizing off-pathway Fe(II) oxidation (Li et al., 2023 []). Cysteine (Cys) acts as both chelator and reductant, forming soluble Fe(III)-complexes that remain active across pH 3–9, increasing the quinolinic acid degradation rate sixfold []. Oxalic acid (OA) operates indirectly by accelerating auxiliary metal cycles—e.g., promoting V(V)/V(IV) cycling in VO2/H2O2 systems—which in turn facilitates Fe(III) reduction [].
Beyond accelerating iron cycling, chemical reductants can modulate reactive species distribution. HA shifts the dominant pathway in Fe(II)/PDS systems from high-valent iron (Fe(IV)) to radical-based routes (SO4/•OH) by rapidly reducing Fe(IV) and regenerating Fe(II) []. In contrast, DTN enhances electron utilization in multi-metal systems via vanadium-mediated cycling, minimizing radical quenching and improving oxidant efficiency [].
In summary, chemical reductants enhance iron-persulfate systems through three core mechanisms: thermodynamic facilitation of Fe(III) reduction, kinetic enhancement of electron transfer, and morphological control of iron speciation. This approach provides a versatile strategy for overcoming the kinetic limitations of iron-based persulfate activation across a wide range of environmental conditions.

3.2.2. Electron Shuttles for Mediated Redox Cycling

The introduction of electron shuttles establishes a reversible electron-transfer system that decouples persulfate activation from the slow kinetics of heterogeneous iron reduction. This strategy employs redox-active mediators capable of cycling between oxidized and reduced states, thereby facilitating continuous Fe(III) reduction and sustaining high catalytic activity without permanent consumption (Figure 5b). In contrast to the static, conductive interfaces created by heteroengineering, electron shuttles operate through dynamic, reversible redox reactions in the bulk solution, mediating electron transfer over distance.
Quinoid compounds function as highly efficient molecular shuttles due to their rapid two-electron redox cycling between quinone (electron-accepting) and hydroquinone (electron-donating) states. In methyl-p-benzoquinone (MBQ)-enhanced systems, this shuttle mechanism operates through precisely coordinated steps: MBQ accepts electrons from electron-rich domains to form methyl-hydroquinone (MHQ), which subsequently reduces Fe(III) to Fe(II). This cyclic electron transfer enables exceptional atrazine degradation across pH 2–7, significantly outperforming conventional Fe(II)/PDS systems []. The universal nature of this mechanism is demonstrated by benzoquinone (BQ), which elevates atrazine degradation from 23.2% to 99.8% in Fe(III)/PMS systems through analogous electron-shuttling pathways.
Natural humic substances represent evolutionarily optimized electron shuttles that integrate quinoid moieties with metal-complexing functional groups. In HA-Fe@BC/PDS systems, the quinone groups in humic acid accept electrons from the conductive biochar matrix to form persistent semiquinone radicals, which drive efficient Fe(III) reduction and enhance rhodamine B degradation kinetics by eightfold []. When coupled with FeAl-layered double hydroxides, dissolved organic matter (DOM) not only facilitates electron transfer but also generates oxygen vacancies that stabilize Fe(II) species. While humic acids mitigate anion interference (Cl/NO3) through complexation, their potential quenching effects at high concentrations necessitate careful dosage optimization.
Sulfur-based compounds introduce advanced electron transfer capabilities through multi-electron redox processes. In the SC-NiFe2S4/PMS system, surface sulfide species (S2−) enable concurrent Ni2+/Ni3+ and Fe2+/Fe3+ cycling while creating oxygen vacancies that synergistically activate PMS through a predominant singlet oxygen (1O2) pathway []. More remarkably, dithiothreitol (DTT) in FeVO4/PMS systems generates strongly reducing SO2 radicals that initiate a cascade electron transfer: reducing V5+ to V4+, which then transfers electrons to Fe(III) through Fe–O–V bonds. This sophisticated electron relay system maintains >90% degradation efficiency over multiple cycles while minimizing metal leaching [].
The strategic integration of electron shuttles represents a paradigm shift in persulfate activation, moving from direct activation to mediated electron transfer regimes. These molecular mediators not only accelerate iron cycling kinetics but also enable unprecedented control over reaction pathways and selectivity, offering a robust strategy for overcoming kinetic limitations in complex environmental matrices.

3.2.3. Elemental Substances as Reductive Modulators

The incorporation of elemental substances—both metallic and nonmetallic—offers a robust strategy to overcome electron-transfer limitations in iron-based persulfate systems. This approach fundamentally alters the local electronic environment and interfacial reaction dynamics, enabling sustained iron cycling through direct electron donation, structural modification, and the creation of synergistic redox partnerships (Figure 5c). This strategy differs from the morphological engineering of catalysts, as it involves the addition of bulk elemental materials (e.g., ZVAl) as co-reagents that corrode to provide electrons, rather than structuring the iron catalyst itself.
Metallic elements such as zero-valent aluminum (ZVAl) and copper function as direct electron donors or co-catalytic mediators. ZVAl and its corrosion products serve dual roles: activating PMS through direct electron transfer to generate SO4, and concentrating pollutants near reactive sites through adsorption, thereby enhancing local reaction efficiency []. Copper enhances iron cycling within bimetallic frameworks (e.g., CuFeBDC); the Cu+/Cu2+ cycle rapidly reduces Fe3+, while Cu+ is regenerated through oxidation by PMS, establishing an autocatalytic loop []. Zinc doping modifies the electronic structure of iron active sites, lowering the redox potential of the Fe3+/Fe2+ couple and increasing electron density, thereby significantly accelerating Fe3+ reduction [].
Nonmetallic elements, particularly boron and carbon, operate through distinct yet complementary mechanisms. Amorphous boron acts as a potent reducing agent at the Fe@NC/B interface, continuously supplying electrons to iron species to drive Fe3+ reduction and potentially directly activating PMS to form both SO4 and 1O2 (Wang et al., 2023 []). Carbon layers simultaneously function as physical barriers (e.g., in carbon-coated nZVI), effectively preventing oxidative passivation of iron nanoparticles and enhancing long-term stability []. Furthermore, specific carbon architectures can induce non-radical pathways, degrading contaminants via direct electron transfer or singlet oxygen generation.
The strategic integration of these elements enables precise control over reaction mechanisms and iron speciation. Elemental additives optimize the local electronic environment of iron active sites—metallic dopants introduce new energy levels and reduce charge transfer resistance, while nonmetallic elements modify work function and surface electron density. These modifications collectively lower the activation energy for PMS decomposition and promote the regeneration of active Fe2+ species. Additionally, elemental enhancers improve catalyst resilience in complex matrices: carbon coatings mitigate interference from common anions (e.g., Cl, CO32−) and natural organic matter, while metallic reductants like ZVAl maintain activity over a broader pH range.
In essence, the use of elemental substances provides a multi-faceted approach to breaking the kinetic constraints of iron cycling. Through electron donation, interface engineering, and structural stabilization, these elements work in concert to sustain efficient persulfate activation, offering a versatile and effective strategy for environmental remediation applications.
In critically evaluating these system optimization strategies, it becomes evident that each approach represents a distinct compromise between electron transfer efficiency and practical implementability. Chemical reductants achieve the most rapid Fe3+ reduction kinetics through direct electron transfer, yet this advantage is counterbalanced by their stoichiometric consumption and potential secondary contamination—fundamental limitations that question their environmental sustainability. Electron shuttles offer an elegant catalytic alternative through reversible redox cycling, but their molecular-level operation renders them vulnerable to matrix interference and degradation, revealing a critical sensitivity to environmental conditions that challenges field application. Elemental substances provide the most robust electron donation platforms through bulk-phase interfaces, yet their macroscopic nature introduces equally macroscopic challenges of metal leaching and surface passivation that ultimately limit longevity. This critical analysis reveals an unresolved tension in the field: the strategies that achieve the fastest iron cycling kinetics tend to be the least sustainable, while the most environmentally compatible approaches often suffer from kinetic limitations. Future research must therefore move beyond optimizing individual strategies and instead focus on developing integrated systems that combine the rapid kinetics of chemical reductants with the sustainability of catalytic mediators and the stability of solid-phase donors, while addressing the fundamental mechanistic trade-offs identified here.
Table 3. Comparative Analysis of System Optimization Strategies via Reductive Enhancement.
Table 3. Comparative Analysis of System Optimization Strategies via Reductive Enhancement.
StrategyCore Mechanism & Action ScalePrimary Role in Iron CyclingKey AdvantageInherent Limitation/RiskReferences
Chemical ReductantsHomogeneous molecular-scale reduction. Provides exogenous electrons via direct donation (e.g., HA, DTN) or complexation-mediated transfer (e.g., AA, Cys).Overcomes kinetic barrier of Fe(III) reduction. Acts as a soluble electron source to rapidly regenerate Fe(II).Immediate efficacy; Broad availability; Can shift reaction pathways (e.g., from Fe(IV) to radicals).Consumable; Requires continuous dosing; Potential formation of harmful byproducts; Increases operational cost.[,]
Electron ShuttlesReversible molecular-mediated transfer. Establishes a catalytic cycle where mediators (e.g., quinones, humics) oscillate between redox states.Decouples Fe(III) reduction from oxidant activation. Enables continuous electron transfer without mediator consumption.Catalytic (non-consumable); High selectivity; Mitigates anion interference; Operates over a wide pH range.Susceptible to radical attack; Performance pH-dependent for natural mediators; Cost barriers for synthetic analogs.[,]
Elemental SubstancesHeterogeneous interface & bulk electron donation. Metallic elements (e.g., ZVAl, Cu0) donate electrons; Carbon matrices act as conductive mediators.Provides a sustained electron flux. Serves as a bulk electron reservoir or creates conductive networks for electron transfer.Sustained release of electrons; Multifunctionality (e.g., adsorption, pH buffering); ZVAl offers a wide pH applicability.Potential metal leaching (e.g., Al3+); Passivation of metal particles over time; Slower initial kinetics compared to molecular reductants.[,]

3.3. External Energy/Field Assistance Strategies

3.3.1. Photo-Assistance for Renewable Electron Injection

Light irradiation serves as a powerful external energy input to overcome the kinetic limitations of iron-mediated persulfate activation by providing a renewable source of electrons and holes (Figure 6a). This strategy fundamentally alters the reaction kinetics and pathways through photogenerated charge carriers that directly drive the critical Fe3+/Fe2+ cycle and facilitate advanced oxidation processes beyond thermal activation (Table 4).
Figure 6. External Field Assistance Strategies. (a) The photocatalytic mechanism. (b) Thermal Activation. (c) Electrocatalytic Assistance. (d) Ultrasound assistance.
The core mechanism involves photonic excitation that generates electron-hole pairs, enabling direct photoreduction of Fe3+ to Fe2+. This process illustrates how visible light irradiation on an Fe/Mn-Prussian blue analog@MXene composite injects photogenerated electrons into the dual metal-cycle pathways. This direct photoreduction bypasses the conventional kinetic barrier of iron reduction, where the corresponding performance data shows the system achieving 93.4% tetracycline degradation []. Beyond accelerating iron cycling, light facilitates the formation of multiple reactive species through both radical and non-radical pathways. In UV/CoFe2O4/periodate systems, photoactivation enables simultaneous generation of six distinct radical species, leading to complete pollutant degradation within 12 min [].
Advanced material design strategies maximize light absorption and charge separation efficiency. Oxygen vacancy-rich ZnFe2O4 (ZFOᵥ) utilizes vacancies as electron reservoirs that localize electrons for persulfate activation, increasing degradation efficiency by 1.23-fold []. The construction of visible-light-responsive heterojunctions (e.g., MoS2/δ-FeOOH) extends absorption spectra while suppressing charge recombination through built-in electric fields, enabling nearly 100% RhB degradation within 35 min []. Defect engineering further enhances performance, as demonstrated by Se vacancies in CoSe2/FeSe2@CNT that modify the electronic structure to favor Fe3+ reduction, increasing radical yield by 3.8 times [].
Photo-assisted systems exhibit remarkable adaptability to complex environmental conditions. The visible light/gallic acid/Fe3+/periodate system maintains high activity across pH 3–9, achieving complete BPA removal in 30 min under neutral conditions []. Natural mineral-based systems such as chalcopyrite/PAA/visible light retain 95% sulfamethoxazole degradation efficiency even in the presence of common anions (Cl/HCO3) []. The integration of photo-catalytic processes with other functions demonstrates multifunctional potential, as shown in FeMn-PBA@MXene systems that simultaneously degrade tetracycline (98.9%) and desalinate water, reducing salinity from 4000 ppm to 240 ppm [].
Both radical and non-radical pathways are enhanced under light irradiation. In CoFe2O4/PMS, SO4 contributes 76% to degradation, though intermediate toxicity necessitates pathway control []. Non-radical routes such as 1O2 generation dominate in certain heterojunctions (e.g., ZnFe2O4/Bi2O2CO3/BiOBr), exceeding the contribution of SO4 []. High-valent iron species (e.g., Fe(IV)=O) also become more accessible under light, showing superior selectivity in contaminant degradation []. DFT simulations provide mechanistic validation, confirming that Fe-modified surfaces significantly reduce PMS adsorption energy and decrease activation barriers [].
In summary, photo-assistance enhances iron-persulfate systems through three fundamental mechanisms: renewable electron injection for Fe3+/Fe2+ cycling, expanded reactive species diversity through photogenerated carriers, and improved environmental adaptability through multi-pathway activation. This approach represents a promising direction for developing sustainable and efficient water treatment technologies that leverage solar energy as a primary driving force.

3.3.2. Thermal Assistance for Overcoming Activation Energy Barriers

Thermal energy activation represents a fundamental strategy for enhancing persulfate-based advanced oxidation processes by providing the necessary energy to overcome intrinsic activation barriers (Figure 6b). This approach operates through two synergistic mechanisms: direct facilitation of O–O bond cleavage in persulfate molecules and thermodynamic optimization of the iron redox cycle, thereby addressing both oxidant activation and catalyst regeneration limitations (Table 4).
The homolytic cleavage of the O–O bond in persulfates requires substantial energy input (Eₐ > 90 kJ/mol). Thermal energy directly supplies this activation energy, enabling efficient persulfate decomposition even in the absence of catalysts. For instance, the thermal activation of PDS for difloxacin degradation exhibits an Eₐ of 91.25 kJ/mol [], significantly higher than that of Fe2+-activated systems (50–70 kJ/mol), underscoring the rate-limiting nature of this step. Beyond direct activation, thermal treatment tunes the iron electronic environment. Vacuum annealing (150–300 °C) increases the Fe2+/Fe3+ ratio in MIL-101(Fe) by up to 50%, thereby optimizing the alignment between the Fe3+/Fe2+ redox potential (E0 ≈ 0.77 V) and that of PMS (HSO5/SO42−, E0 = 1.82 V). This electronic optimization enhances the electron transfer rate constant by 2.3 times [], significantly accelerating the iron-mediated activation cycle. However, this technique also presents significant limitations including equipment complexity, high energy consumption for maintaining vacuum conditions, scalability challenges for industrial applications, potential structural damage from prolonged treatment, and limitation to solid catalyst systems only [].
The iron-mediated activation follows a chain reaction mechanism where thermal energy alleviates the kinetic bottleneck through two complementary pathways. First, it reduces the activation energy of Fe3+ reduction—as demonstrated in a coal/Fe2O3 system where Eₐ decreased from 62.79 kJ/mol to 56.50 kJ/mol at 60 °C, increasing the reduction rate threefold []. Second, autothermal reinforcement establishes a self-sustaining cycle whereby the exothermic nature of the Fe2+-PMS reaction (ΔH = −198 kJ/mol) maintains elevated temperature and promotes continuous radical generation, boosting radical yield by 3–5 times compared to non-thermal systems. The dominance of SO4 (>70%) under such thermal activation has been confirmed via ESR in systems including NO removal and H2S oxidation [].
The integration of solar energy introduces a renewable pathway for thermal activation, combining photonic and thermal effects for enhanced efficiency. Solar irradiation provides both electronic excitation and thermal energy, enabling multi-field synergies. For example, UV-induced ligand-to-metal charge transfer (LMCT) in Fe3+-carboxylate complexes enhances the quantum yield of Fe3+ reduction, while the thermal component (>40 °C) facilitates PMS activation. In a solar/Fe2+/PMS system, the E. coli inactivation rate constant reaches 0.23 min−1—10 times higher than photolysis alone []. Low-grade solar-thermal energy (∼60 °C) can efficiently activate PDS for antibiotic degradation, even overcoming an Eₐ of ∼91.25 kJ/mol [].
Thermal assistance demonstrates remarkable adaptability to various catalyst architectures and environmental conditions. The maintained elevated temperature prevents iron precipitation and stabilizes reactive intermediates across a broader pH range than conventional systems. Furthermore, the reduced reliance on precise catalyst design makes thermal activation particularly suitable for complex wastewater matrices where catalyst poisoning typically occurs.
In essence, thermal enhancement operates through dual fundamental mechanisms: thermodynamic facilitation of persulfate bond cleavage and kinetic acceleration of the iron redox cycle. This approach provides a versatile and often essential strategy for achieving efficient pollutant degradation, particularly for recalcitrant compounds requiring high activation energies. The integration with renewable solar-thermal energy further positions thermal activation as a sustainable pathway for advanced oxidation processes.

3.3.3. Electrocatalytic Assistance for Controlled Iron Redox Cycling

Electrocatalytic activation represents a sophisticated strategy to overcome the kinetic limitations of iron-based persulfate systems through precise electronic control (Figure 6c). By applying an external potential, this method enables directional electron transfer and dynamic regulation of iron speciation, achieving sustained Fe2+ regeneration and significantly enhanced catalytic efficiency without chemical consumables (Table 4).
The fundamental advantage of electrochemical assistance lies in its ability to promote the cathodic reduction of Fe3+ to Fe2+, thereby breaking the rate-limiting step in traditional Fenton-like systems. In electro/Fe3+/PDS systems, cathodic electron transfer efficiently regenerates Fe2+, which activates persulfate to generate sulfate radicals (SO4) and hydroxyl radicals (•OH), enabling effective degradation of recalcitrant pollutants such as clofibric acid []. Heterogeneous catalysts like Fe-B further enhance the electron transfer kinetics between Fe3+ and Fe2+ under an electric field, considerably improving PMS activation efficiency []. This electrochemical acceleration addresses the intrinsic slow kinetics that plague conventional iron cycling systems.
Different iron species exhibit distinct activation behaviors under electrochemical promotion. Zero-valent iron (ZVI) serves dual roles as both a Fe2+ source and electron donor in ZVI-E-Fenton-PMS systems: ZVI undergoes anodic oxidative dissolution to release Fe2+, while the cathode simultaneously reduces Fe3+, creating a synergistic effect that promotes the formation of multiple radical species including •OH, SO4, and 1O2 []. For crystalline iron minerals such as pyrite (FeS2), electrochemical assistance leverages surface sulfide (S2−) as an electron mediator to facilitate Fe2+ regeneration. Simultaneously, sulfur intermediates (e.g., Sₙ2−, S0) enhance radical production, revealing a novel mineral-electric field synergistic mechanism [].
Advanced electrode and catalyst designs have optimized iron cycling efficiency and system sustainability. Self-sustaining cathode materials (e.g., Fe3O4-CaO2) utilize the chemical cycle of CaO2/Ca(OH)2 to accelerate Fe3+ reduction, reducing energy consumption to 5% of that in conventional systems []. The introduction of atomic hydrogen (H*) via Pd/Al2O3 particle electrodes offers an efficient pathway for reducing Fe3+ while significantly minimizing iron sludge formation []. These systems perform effectively under near-neutral conditions (pH ~6.0) via heterogeneous iron (oxyhydr)oxides, overcoming the strong acidity requirement of traditional Fenton reactions [].
Electrocatalytic activation provides exceptional control over reaction mechanisms, enabling a deliberate shift from radical to non-radical pathways to reduce scavenging effects and improve selectivity. In the E/N-S@Fe-PBC/PMS system, the combination of an electric field and nitrogen-sulfur co-doped biochar significantly promotes the generation of singlet oxygen (1O2), contributing up to 37% to pollutant degradation []. Electrode material selection profoundly influences the activation route: BDD anodes favor direct electron transfer to produce SO4, while DSA anodes promote surface-mediated non-radical oxidation []. This tunability provides a strategic advantage for treating complex matrices with varying compositions.
Despite these promising features, challenges remain in electrode cost, interference from co-existing substances, and long-term operational stability. Future research should focus on developing robust and affordable electrode materials, elucidating interfacial iron cycling mechanisms through in situ techniques, and demonstrating scalability in real-world environments. The integration of electrocatalytic processes with renewable energy sources further represents a critical pathway toward sustainable water treatment applications.

3.3.4. Ultrasonic Assistance for Enhanced Interfacial Processes

Ultrasound irradiation enhances iron-based persulfate activation through cavitation-induced physical and chemical effects that fundamentally alter interfacial processes (Figure 6d). This approach operates through three primary mechanisms: microjet impingement that removes passivation layers, enhanced mass transfer that accelerates reactant availability, and sonochemical generation of reactive species that complement persulfate activation (Table 4).
The collapse of cavitation bubbles generates extreme local conditions (~5000 K, ~1000 bar) and powerful microjets that mechanically disrupt passivating iron oxide layers. In zero-valent iron (ZVI) systems, this continuous surface renewal facilitates the corrosion of Fe0 to release Fe2+, which activates persulfates to generate SO4 []. This process significantly improves electron transfer from Fe0 to PMS, enabling rapid degradation—99.76% removal of RhB within 12 min—while simultaneously promoting the reduction of Fe3+ back to Fe2+ []. The mechanical energy input alleviates Fe3+ accumulation and prevents reaction stagnation, overcoming a fundamental limitation of conventional homogeneous systems.
In heterogeneous catalytic systems, ultrasound enhances iron cycling through synergistic redox mechanisms. In FeS-based catalysts, S2− oxidation to polysulfides (Sx2−) releases electrons that drive Fe3+ reduction to Fe2+. The regenerated Fe2+ then activates persulfate to produce radicals, establishing a self-sustaining cycle []. This mechanism remains highly efficient under neutral conditions, achieving 86.6% degradation of diclofenac and overcoming the pH limitations of homogeneous Fenton systems []. Material architecture plays a crucial role: porous sponge iron (Fe1), with its high surface area, provides abundant Fe2+/Fe3+ sites and exhibits higher iron transformation rates than powdered iron (Fe0), significantly enhancing sludge dewatering (reducing CST by 89%) and soluble organic matter degradation [].
Ultrasound assistance maintains long-term catalytic activity by countering deactivation processes. Cavitation microjets continuously remove passive oxide layers from material surfaces such as Fe3O4 or ZVI, exposing fresh active sites []. Support materials contribute to stability; in Fe3O4@biomass porous carbon (BPC), the carbon carrier disperses iron particles, reduces agglomeration, and preserves reactive interfaces, maintaining 70% norfloxacin degradation after four cycles []. Additionally, ultrasound induces mechanochemical effects that rearrange surface metal valences. For example, in Fe-NiOx catalysts, sonication exposes more high-valence metal sites (Ni3+/Fe4+), thereby improving PMS activation efficiency [].
The integration of ultrasound with persulfate activation demonstrates particular effectiveness in complex matrices where conventional systems fail. The physical cleaning action mitigates fouling from natural organic matter, while the enhanced mixing overcomes diffusion limitations in high-viscosity environments. Furthermore, the simultaneous production of hydroxyl radicals from water sonolysis creates complementary oxidation pathways that enhance overall treatment efficiency.
Ultrasonic assistance enhances iron-persulfate systems through three convergent mechanisms: surface renewal via cavitational cleaning, enhanced mass and electron transfer through turbulent mixing, and sonochemical radical generation. This multi-mechanism approach provides a robust strategy for maintaining catalytic activity, preventing deactivation, and enabling operation under environmentally relevant conditions.
In critically assessing external field-assisted iron cycling, fundamental distinctions emerge in how different energy inputs activate and sustain the Fe3+/Fe2+ redox cycle. Photo-assistance drives charge separation and electron injection through photon absorption, enabling renewable Fe3+ reduction but facing intrinsic limitations in light penetration and quantum efficiency [,]. Thermal activation provides the requisite energy for persulfate homolysis and accelerates electron transfer kinetics, yet the substantial energy input raises concerns for sustainable implementation [,]. Electrochemical systems achieve precise control over iron speciation through applied potentials, enabling directed electron transfer without chemical additives but confronting challenges of electrode stability and mass transport limitations [,]. Ultrasonic activation uniquely combines physical cleaning of passivated surfaces with enhanced mass transfer through cavitation effects, though energy intensity and reactor scalability remain significant barriers [,]. Collectively, these approaches reveal a critical trade-off: external energy inputs can dramatically enhance iron cycling kinetics but often at the cost of operational complexity and energy consumption that may limit practical application. Future advances should focus on developing energy-adaptive systems that optimize the synergy between multiple activation mechanisms while minimizing external energy requirements through intelligent catalyst design and process integration.
Table 4. Comparative Analysis of External Field Assistance Strategies for Iron Cycling Enhancement.
Table 4. Comparative Analysis of External Field Assistance Strategies for Iron Cycling Enhancement.
StrategyFundamental Principle & Action ScalePrimary Role in
Iron Cycling
Unique AdvantagePractical Consideration & LimitationReferences
Photo-Assistance Photon-electron conversion. Utilizes photonic energy to generate electron-hole pairs for Fe3+ photoreduction and radical generation.Provides renewable electrons; Expands reactive species diversity (e.g., via LMCT).Solar energy utilization; Enables multi-pathway activation; High tunability via material design.Dependent on light penetration; Catalyst requires photoactivity; Possible light shielding in complex matrices.[,]
Thermal Assistance Thermal energy input. Overcomes activation energy barriers for O–O bond cleavage and accelerates reaction kinetics.Lowers activation energy for both persulfate decomposition and Fe3+ reduction; Enables autothermal cycles.Universal applicability; No need for specialized catalysts; Effective for recalcitrant compounds.High energy consumption; Limited control over reaction pathways; Potential for byproduct formation.[,]
Electrocatalytic Assistance Precise electron delivery. Applies potential to directionally drive electrons to Fe3+ at the cathode interface.Sustains controlled Fe2+ regeneration via direct cathodic reduction; Decouples oxidation and reduction sites.Precise redox control; Tunable reaction mechanisms; Minimal chemical consumption.Electrode cost and fouling; Mass transfer limitations; System complexity and scalability.[,]
Ultrasonic AssistanceCavitation-induced phenomena. Uses acoustic cavitation for surface cleaning, enhanced mass transfer, and sonochemistry. Prevents passivation via surface renewal; Enhances mass transfer of reactants to active sites.Operates under any pH; Mitigates catalyst fouling; Effective in viscous or particulate-laden systems.High energy intensity; Limited reactor design scalability; Potential for equipment erosion.[,]

4. Application and Performance Evaluation in Environmental Remediation

4.1. Mechanism-Tailored Strategies for Different Contaminant Classes

The deployment of iron-based persulfate systems for environmental remediation necessitates a mechanistic understanding of how enhanced iron cycling translates into effective pollutant degradation across diverse contaminant classes. Performance evaluation must extend beyond simple degradation metrics to encompass pathway selectivity, matrix effects, and long-term sustainability, establishing a critical link between fundamental mechanisms and practical application.
Molecular architecture dictates contaminant susceptibility to specific oxidative pathways, requiring strategic alignment between iron cycling enhancement and target pollutant characteristics. Electron-rich compounds, including phenolics and dyes, undergo preferential degradation through non-radical pathways (e.g., direct electron transfer or singlet oxygen), as their high electron density favors selective oxidation (Figure 7a). Crucially, the oxidative pathway shifts for contaminants where the primary challenge is not initial electron abstraction, but the cleavage of inherently stable molecular frameworks. However, for refractory aromatic compounds, the sustained generation of SO4 and •OH proves essential for forceful aromatic ring cleavage and complete antibiotic inactivation []. For electron-withdrawing contaminants such as benzoic acid and nitrobenzene, non-radical pathways (e.g., 1O2 or surface electron transfer) often exhibit limited efficacy. This is primarily due to the reduced electron density of the pollutant’s aromatic ring, which hinders its interaction with mild oxidants. Bimetallic systems establish coupled redox cycles that bypass kinetic limitations of solitary iron cycling: in Fe/Mo@C, Mo doping stabilizes lower iron valence states while modulating the d-band center of Fe active sites, enhancing PDS activation through facilitated O–O bond cleavage and achieving a 26-fold increase in sulfamethoxazole degradation rate []. Halogenated contaminants present distinct challenges requiring specialized approaches. While pesticides undergo radical-mediated dehalogenation, Halogenated Substituted Benzene resist conventional oxidation due to exceptional C–X bond stability (Figure 7c). Sulfur-mediated strategies address this dichotomy through enhanced iron cycling; surface S2− species in FeS@biochar provide electrons for Fe(III) reduction, doubling iron turnover rates for 95.9% 2,4-D degradation [].
Figure 7. Mechanism-Tailored Strategies for Different Contaminant Classes. (a) The Electron-rich compound degradation. (b) The Refractory aromatic compounds. (c) The halogenated contaminants. (d) The Heavy metal–organic complexes.
While conventional radical pathways often prove inadequate for PFAS degradation due to radical scavenging and the exceptional stability of C-F bonds, iron-activated persulfate systems achieve effective PFOA removal through integrated mechanisms. Research demonstrates that zero-valent iron (ZVI) synergistically activates persulfate by generating sulfate radicals while simultaneously facilitating PFOA decomposition, achieving 67.6% degradation at 90 °C []. More significantly, heterogeneous systems employing FeS overcome the limitations of homogeneous radical reactions by concentrating PFOA at the mineral surface (KD = 100–500 L/kg), creating a localized reaction environment that minimizes interference from common water constituents []. This surface-mediated approach enables efficient PFOA degradation even in complex water matrices. Furthermore, ferrous iron activation provides a practical pathway for ambient temperature treatment, achieving 64% PFOA degradation within 4 h under anoxic conditions []. These iron-based systems thus address PFAS persistence through combined surface enrichment, localized radical generation, and adaptable activation conditions that collectively enhance degradation efficiency.
The treatment of heavy metal–organic complexes (e.g., metal-EDTA, metal-citrate) presents a unique challenge that requires sequential processing: initial decomplexation to break the stable coordination spheres, followed by immobilization of the liberated metal ions (Figure 7d). Enhanced iron cycling provides an elegant solution to this two-stage problem through sustained generation of reactive species and facilitated electron transfer processes. The decomplexation stage primarily relies on radical-mediated oxidation, where accelerated Fe(II)/Fe(III) cycling in systems such as electrochemical Fe-MOF/persulfate enables continuous production of hydroxyl radicals (•OH) []. These strongly oxidizing species selectively attack the coordination bonds in complexes like Cu-EDTA through hydrogen abstraction and ligand oxidation pathways, effectively liberating free Cu2+ ions while partially mineralizing the organic ligands. The subsequent immobilization stage then addresses the fate of these freed metal ions through multiple mechanisms: in electrochemical systems, the applied potential drives cathodic reduction and deposition of metal ions, effectively recovering them in elemental form; in parallel, the iron-based catalysts themselves provide adsorption sites and nucleation centers for metal hydroxide precipitation. This integrated approach is further exemplified in Fe(II)-Al layered double hydroxides, where structural Fe–OH sites serve as electron transfer bridges, facilitating both the oxidation of arsenite (As(III)) to arsenate (As(V)) and subsequent adsorption of the transformed arsenic species through surface complexation and structural incorporation []. The synergy between enhanced iron cycling and these immobilization pathways ensures not only effective decomplexation but also prevents re-complexation and secondary contamination, offering a comprehensive solution for treating persistent metal–organic complexes that surpasses conventional treatment limitations.
In conclusion, the strategic enhancement of iron cycling is not a universal remedy but a versatile toolkit. The key to precision remediation lies in diagnosing the contaminant’s molecular identity and matching it with a tailored iron cycling strategy—be it to generate powerful radicals, selective non-radical oxidants, or to enable coupled degradation-immobilization processes. This mechanism-based framework provides a rational design principle for developing targeted, efficient, and sustainable water treatment technologies.

4.2. Adaptability to Complex Environmental Matrices

The performance of iron-based persulfate systems in real-world applications is critically challenged by complex environmental matrices. Background constituents, such as inorganic anions and natural organic matter (NOM), can scavenge reactive species, compete for active sites, and alter reaction pathways. Strategies that enhance iron cycling demonstrate superior resilience to these interfering species by sustaining catalytic activity through robust electron transfer and surface-mediated radical generation, rather than relying solely on short-lived radicals in bulk solution.
The presence of common anions (e.g., Cl, HCO3, H2PO4) typically quenches free radicals and suppresses degradation efficiency. However, systems with reinforced iron cycling maintain significant activity through enhanced interfacial processes. The magnetic FeS@biochar (MFB-500)/PMS system exhibits strong resilience toward Cl, NO3, and SO42− (<20% inhibition in 2,4-D degradation), though performance declines with HCO3 and H2PO4 due to their radical quenching and precipitation effects []. The biochar matrix facilitates electron transfer and surface Fe(II) regeneration, thereby reducing dependence on free radicals and enhancing tolerance to anionic interference. Similarly, Sr-doped LaFeO3 (LSFO-40) retains over 90% degradation efficiency of Orange I under coexisting Cl and HCO3 across pH 5–9. Doping increases oxygen vacancy concentration, which accelerates Fe(III)/Fe(II) turnover and strengthens PMS adsorption, effectively minimizing competitive anion binding [].
Natural organic matter such as humic acid (HA) can inhibit degradation by blocking active sites or scavenging reactive species. Nevertheless, systems with enhanced iron cycling demonstrate notable resistance. Magnetic MnFe2O4 synthesized from spent lithium-ion batteries (MFO-LIBs) achieves 85% bisphenol A degradation even under high HA concentration (20 mg/L) and alkaline conditions (pH 10). The heterostructure provides multiple redox couples (Mn(II)/Mn(III) and Fe(II)/Fe(III)) that maintain catalytic activity with limited influence from HA []. In another example, the presence of HA causes only an 8% decrease in bisphenol A degradation efficiency in a CoFe2O4-LIBs/PMS system. The highly synergistic Co(II)/Co(III) and Fe(II)/Fe(III) cycles compensated for radical consumption by HA, underscoring the role of continuous metal redox cycling in mitigating NOM-induced inhibition [].
Real wastewater systems introduce additional complexity, including heavy metals and oily constituents, which can further challenge treatment efficiency. In an nZVI-PS system applied to e-waste wastewater, coexisting Cu2+ promotes a cooperative redox cycle involving Cu0–Cu(II) and Fe(II)/Fe(III), which enhances persulfate activation and increases BDE-47 degradation by 40%. In contrast, Zn2+ and Ni2+ adsorb onto catalytic surfaces and inhibit reaction efficiency []. Similarly, oil-sludge carbon-supported CoFe2O4 (CoFe2O4/OSC) achieves 90.8% norfloxacin degradation in an oily water matrix. The dual metal cycling between Co and Fe counteracts scavenging by organic constituents, while metal leaching remains low (<30 μg/L), illustrating the potential of iron-enhanced systems under challenging environmental conditions [].
In summary, the adaptability of iron-persulfate systems to complex matrices is significantly improved by strategies that enhance iron cycling and electron transfer mechanisms. By reducing reliance on highly scavengeable radicals and promoting surface-mediated or direct electron transfer processes, these advanced systems sustain oxidation capacity and support practical application in realistic water and wastewater treatment scenarios.

4.3. Pollutant Degradation Kinetics and Mineralization Efficiency

The enhancement of iron cycling efficiency directly translates to superior performance in pollutant degradation, which is quantitatively analyzed through reaction kinetics and mineralization efficiency. The degradation kinetics in these complex systems are most commonly fitted with a pseudo-first-order model (−ln(C/C0) = kobs t), where the observed rate constant (kobs) serves as a key metric for the net efficiency of persulfate activation and pollutant oxidation. This model is applicable under the common condition of excess persulfate, where the kinetics are governed by the catalyst-mediated generation of reactive species and their subsequent reaction with the target pollutant. The strategies discussed below enhance kobs by fundamentally accelerating the rate-limiting Fe3+/Fe2+ cycle. Furthermore, the ultimate remediation goal is mineralization, measured as Total Organic Carbon (TOC) removal, which reflects the complete oxidation of organic pollutants to CO2 and water, a process often lagging behind initial pollutant degradation due to the formation of persistent intermediates.
Reductants such as sulfides and polyphenols significantly accelerate the Fe(III)/Fe(II) cycle, providing a continuous supply of Fe2+ for persulfate activation. Sulfur-modified ZVI (S-mZVI) activated PMS to degrade sulfamethoxazole (SMX) with an observed rate constant (kobs) of 0.1403 min−1, representing a 29.4% increase over the unmodified system (Li et al., 2019 []). The addition of gallic acid (GA) to a Fe3+/PMS system increased the degradation rate constant of BDE47 by 9.4 times (kobs = 0.298 h−1), attributable to the GA oxidation intermediates that continuously regenerate Fe2+ from Fe3+ []. The increase in kobs is directly tied to the accelerated Fe3+ reduction kinetics provided by these reductants. However, the mineralization efficiency is highly dependent on the nature of the reductant and the resultant reactive species; for instance, reductants that promote a radical-dominated pathway (e.g., SO4 and •OH) typically achieve more complete TOC removal than those favoring selective non-radical pathways.
Bimetallic systems establish complementary redox couples that facilitate electron transfer, markedly improving iron cycling kinetics. In a Cu0/Fe3O4-PMS system, the degradation rate of Rhodamine B increased 36-fold compared to Fe3O4 alone, due to Cu0-mediated reduction of Fe3+ and enhanced singlet oxygen production []. Pd/Al2O3 activated PMS for degrading 1,4-dioxane with a metal-loading normalized pseudo-first-order rate constant 16,800 times higher than that of a Cu-Fe bimetallic system, as Pd nanoparticles efficiently recycled surface Fe2+ []. This dramatic increase in kobs stems from the establishment of a synergistic redox cycle that lowers the activation energy for the critical Fe2+ regeneration step. The continuous and concurrent generation of multiple radical species (e.g., SO4, •OH) in these systems is particularly effective for fragmenting aromatic rings and oxidizing intermediate products, thereby bridging the gap between rapid degradation and high mineralization efficiency.
Carbon materials serve as electron mediators that enhance iron cycling through conductive networks and defect-induced catalysis. Graphitized nanodiamond (G-ND) activated PMS for phenolic degradation, achieving a TOC removal >65%, as its sp2-carbon surface facilitated direct electron transfer from pollutants to persulfate []. Encapsulated nZVI@biochar (BC-Fe0) activated PDS to achieve 100% removal of paracetamol within 20 min (kobs = 0.37475 min−1), where the carbon shell prevented Fe0 passivation and promoted non-radical pathways [].
External energy fields such as weak magnetic fields (WMFs) and light irradiation overcome diffusion and kinetic barriers in iron-activated systems. The application of a WMF to a Fe0/PMS system increased the kobs for SMX degradation by 2–3 times, owing to magnetically enhanced convection that accelerated Fe0 corrosion and Fe2+ release []. While effective, these approaches often require additional infrastructure and energy input, and their effectiveness may be limited for non-magnetic or homogeneous catalysts.
In summary, the enhancement of iron cycling unequivocally accelerates pollutant degradation kinetics (kobs) by overcoming the electron-transfer bottleneck. More profoundly, the pathway to efficient mineralization is dictated by the nature of the dominant oxidizing species sustained by these strategies. Systems that foster a high and continuous flux of strong radicals, particularly SO4 and •OH, generally achieve superior TOC removal. This is because these non-selective radicals are uniquely capable of unselectively attacking and fragmenting the diverse and recalcitrant intermediate compounds that accumulate during oxidation. In contrast, while highly selective, non-radical pathways (e.g., 1O2, direct electron transfer) often exhibit lower mineralization potentials. Therefore, the rational selection of an iron-cycle enhancement strategy must balance the goal of rapid initial degradation (high kobs) with the requirement for deep oxidation (high TOC removal), by deliberately steering the reaction pathway towards radical generation where complete mineralization is the priority.

4.4. Catalyst Stability and Recyclability in Iron-Enhanced Persulfate Systems

The transition from laboratory research to practical application of iron-based persulfate systems critically depends on catalyst stability and recyclability. Enhanced iron cycling strategies directly address the fundamental challenges of active site preservation and structural integrity maintenance over extended operational periods [,]. These advancements primarily operate through the introduction of auxiliary redox couples, strategic physical confinement of active components, and the innovative design of self-regenerative interfaces—all contributing to sustained catalytic performance.
Multivalent elemental doping significantly improves stability by establishing complementary redox systems that reduce dependence on iron dissolution. In CuFe2O4-CoFe2O4 composites, synergistic interactions between Co2+/Co3+ and Cu+/Cu2+ cycles promote efficient Fe2+/Fe3+ cycling while markedly reducing metal leaching, maintaining 94.3% degradation efficiency after five operational cycles []. The stabilization of coordinatively unsaturated iron sites in mixed-valence systems similarly demonstrates how alternative redox pathways can alleviate irreversible iron loss, though potential environmental implications of secondary metal leaching require careful consideration [].
Structural confinement strategies effectively immobilize active iron species through tailored support interactions and engineered interfaces. Encapsulation of iron oxides within carbonaceous matrices suppresses metal leaching via strong electronic interactions, with some systems demonstrating iron loss rates below 0.125 wt% while retaining over 87% catalytic activity after multiple cycles []. Biochar-supported systems utilize both dispersive effects and surface functional groups to stabilize iron species, achieving minimal leaching through combined physical and chemical stabilization mechanisms []. The critical importance of interfacial bonding strength is increasingly recognized, as weaker interactions may lead to structural degradation under challenging operational conditions [].
Advanced self-regenerative systems represent a paradigm shift in catalyst design, where cocatalysts enable continuous iron cycling through interfacial electron transfer. Metal sulfides such as MoS2 function as effective electron mediators, reducing Fe3+ to Fe2+ while undergoing controlled oxidation themselves, decreasing iron leaching rates to 0.33% of initial levels while maintaining >80% efficiency over six cycles []. WS2 as a cocatalyst enhances electron transfer via surface sulfur vacancies, improving iron cycle efficiency tenfold []. The long-term efficacy of these systems depends critically on the chemical stability of the cocatalyst components, as structural transformations under operational conditions may gradually diminish their electron transfer capability [].
The regeneration of spent catalysts is a key factor for long-term operation. Common regeneration methods include simple water washing to remove surface-adsorbed species, thermal treatment (e.g., annealing at 300–500 °C in inert atmosphere) to restore surface properties and remove carbonaceous deposits, and chemical washing with solvents or mild oxidants. However, catalysts invariably experience varying degrees of activity loss upon regeneration, primarily due to irreversible structural changes, metal leaching accumulation, and the inability to fully restore the original active site coordination. The reported number of effective regeneration cycles in the literature is typically in the range of 3 to 6 cycles while maintaining >80% of the initial activity, as exemplified by systems like Fe/Mo@C [] and FeCo@NBC []. More robust designs, such as those employing strong structural confinement [], demonstrate the potential for extended stability over even more cycles, though a gradual decline in performance is generally observed due to the aforementioned irreversible changes.
Despite these significant advancements, the field requires more rigorous validation under environmentally relevant conditions. Most current studies report only short-term cycling performance, leaving open questions about long-term stability in complex aqueous matrices containing high salinity or organic loadings []. Furthermore, standardization of leaching assessment protocols is urgently needed, as current methodologies vary widely in their detection limits and speciation capabilities. Future research should prioritize in situ characterization of iron valence states and structural evolution during extended operation, providing deeper insights into deactivation mechanisms and regeneration possibilities.
In conclusion, enhancing iron cycling through multivalent doping, structural confinement, and regenerative interfaces provides a multifaceted strategy for improving catalyst stability and recyclability. The continued advancement of these systems toward practical implementation will require combined efforts in materials design, operational optimization, and standardized stability assessment under realistic environmental conditions.

4.5. Environmental Risk and Cost-Effectiveness Analysis of Iron-Cycle Enhanced Systems

While iron-cycle enhancement strategies significantly improve the catalytic performance of persulfate-based advanced oxidation processes, their practical implementation requires careful evaluation of environmental risks and economic feasibility. A comprehensive analysis reveals several critical challenges that must be addressed before large-scale application.
Environmental risk considerations primarily focus on metal leaching and transformation byproduct formation. Although carrier immobilization strategies such as biochar-supported Fe3O4 can reduce iron leaching to below 0.125 wt% after multiple cycles, bimetallic systems introduce additional concerns regarding secondary metal leaching (e.g., Co, Cu) that may cause secondary pollution []. More importantly, the alteration of degradation pathways due to enhanced iron cycling may generate more toxic intermediates than the parent compounds. Radical-dominated reactions tend to produce hazardous chlorinated quinones and nitro-byproducts during antibiotic degradation, while non-radical pathways, though potentially reducing halogenated byproducts, rely on the stability of high-valent metal species whose fluctuation may increase treatment uncertainty [,,]. Furthermore, iron precipitates such as Fe(OH)3 flocs can adsorb toxic transformation products, forming “iron-toxin” co-precipitates that accumulate in sediments and pose potential ecological risks [].
Economic feasibility analysis indicates that cost considerations extend beyond simple catalyst synthesis to encompass operational expenses and long-term sustainability. High-performance catalysts including Fe/Mo-CNs and Sr-doped LaFeO3 require complex preparation procedures that increase material costs [,]. Electrochemically assisted strategies, while effective in accelerating iron cycling, consume substantial energy (up to 2.51 kWh/m3), significantly exceeding the costs of conventional Fenton systems []. The continuous addition of chemical reductants (e.g., hydroxylamine, cysteine) further increases operational costs and raises questions about long-term economic viability [,].
Promising strategies for enhanced sustainability emerge through waste resource utilization and renewable energy integration. Catalysts derived from waste lithium batteries or sludge-derived biochar not only reduce raw material costs by 30–50% but also enable value-added use of solid waste [,]. The incorporation of crystalline boron, despite increasing reaction rates by 65-fold, raises concerns about potential water eutrophication due to boron residue, highlighting the need for comprehensive environmental risk assessment even for effective enhancers []. Similarly, while MoO2 modification effectively promotes Fe2+/Fe3+ cycling, the long-term ecotoxicity of molybdenum remains insufficiently studied, requiring further investigation into its environmental fate and impact [].
In summary, while iron-cycle enhancement strategies show remarkable improvements in catalytic performance, their practical implementation necessitates a balanced consideration of environmental risks and economic factors. Future advancements in this field should focus on developing comprehensive assessment protocols that simultaneously address treatment efficiency, economic viability, and environmental safety to ensure the sustainable development of iron-persulfate systems for water treatment applications.

4.6. Comparative Analysis with Conventional AOPs

A systematic comparison reveals both advantages and limitations of iron redox cycling-based persulfate processes relative to established AOPs. In terms of degradation efficiency, these systems generate multiple reactive species including sulfate radicals (SO4, E0 = 2.5–3.1 V), hydroxyl radicals (•OH), and high-valent iron species, enabling effective degradation of recalcitrant compounds that resist conventional •OH-based oxidation [,,]. The SO4 demonstrates particular effectiveness against electron-rich organic contaminants due to its selective oxidation mechanism and longer half-life compared to •OH [,].
Operationally, iron-persulfate systems function effectively across pH 3–9, significantly broader than classical Fenton’s narrow optimal range (pH 2.5–3.5) where iron precipitation becomes problematic [,,]. While homogeneous Fe2+/persulfate systems operate with minimal energy input, the enhanced systems employing thermal activation (60–80 °C) or electrochemical assistance incur substantial energy costs that may approach those of UV/H2O2 processes [,].
Environmental considerations present distinct challenges. Unlike Fenton processes that generate substantial iron sludge (often 10–30% of treated water volume), properly designed heterogeneous systems significantly reduce solid waste [,]. However, sulfate accumulation from persulfate decomposition (typically 200–500 mg/L SO42− per 1 g/L persulfate dose) represents a unique concern absent in H2O2-based systems []. Furthermore, advanced bimetallic catalysts, while enhancing iron cycling efficiency, introduce risks of secondary metal leaching (e.g., Co, Cu) that require careful management [,].
From a practical perspective, these systems show particular promise for in situ remediation scenarios where persulfate’s subsurface stability (half-life of weeks to months) provides significant advantages over the rapid decomposition of H2O2 (half-life of hours) [,]. They also demonstrate superior performance for treating halogenated compounds and certain pharmaceuticals where SO4 exhibits higher oxidation efficiency than •OH [,]. While full-scale implementation currently lags behind established technologies like ozonation and conventional Fenton, ongoing advances in catalyst design and reactor engineering are rapidly closing this gap.
This analysis confirms that iron redox cycling-based persulfate processes occupy a valuable niche in the AOP spectrum, particularly suited for challenging remediation scenarios where their multi-oxidant capacity and operational flexibility provide distinct advantages, provided that sulfate accumulation and potential metal leaching are properly managed.

5. Challenges and Future Perspectives

The advancement of iron redox cycling for persulfate activation presents a complex interplay between promising laboratory demonstrations and the multifaceted realities of environmental application. This section critically examines the current challenges rooted in mechanistic, material, and system-level constraints, and reframes future research needs as concrete opportunities to bridge the gap between fundamental insight and sustainable implementation.

5.1. Current Research Challenges

While numerous studies have demonstrated enhanced performance through various strategies, the precise mechanisms governing interfacial electron transfer and radical/non-radical pathway selection remain inadequately resolved. This complexity, arising from the generation of multiple reactive species (SO4, •OH, 1O2, Fe(IV)=O), can offer a broader reactivity profile against recalcitrant pollutants compared to the single-radical •OH-dominated chemistry of conventional Fenton systems [], but it concurrently complicates mechanistic elucidation and control. The complexity of multiphase systems, where homogeneous and heterogeneous processes coexist, creates significant challenges in decoupling individual contribution mechanisms. The lack of standardized protocols for identifying reactive oxygen species further complicates cross-study comparisons and mechanistic validation.
The current technological readiness level of most advanced iron-persulfate systems remains low, with complex catalysts like single-atom Fe-N-C sites and core–shell heterostructures typically demonstrating only single-digit operational cycles in laboratory studies [,,]. This represents a significant gap from the >50 cycles required for practical water treatment applications. Many high-performance catalysts rely on precious or toxic metals (e.g., Co, Mo, Pd) and energy-intensive synthesis methods, raising concerns about environmental footprint and scalability. This reality contrasts sharply with the ideal of sustainable environmental remediation. Economic viability analysis reveals substantial challenges, with electrochemical assistance consuming significantly more energy than conventional treatment [], while continuous reductant addition increases operational costs []. The synthesis of complex nanostructures often involves multiple steps, organic solvents, and high-temperature treatments that are difficult to scale economically. Moreover, the long-term stability of these advanced materials under realistic environmental conditions remains largely unverified, with most studies reporting fewer than ten operational cycles under idealized laboratory conditions.
A pronounced performance gap exists between laboratory and field conditions, where degradation efficiency typically decreases significantly in real wastewater due to radical scavenging by background constituents [,]. Similarly, iron leaching rates increase substantially when transitioning from ultrapure water to real wastewater matrices [,]. The transition from batch-scale experiments to continuous-flow systems presents substantial engineering hurdles. Mass transfer limitations, catalyst recovery and reuse, and energy input requirements become critical factors at larger scales. Additionally, the presence of complex matrices in real wastewater—including suspended solids, natural organic matter, and competing ions—can significantly diminish system performance through catalyst fouling, radical scavenging, and active site blockage.
Current research predominantly focuses on performance metrics while paying insufficient attention to potential secondary pollution risks. The leaching of metal ions (both iron and dopants), formation of toxic transformation products, and long-term ecotoxicity of catalyst components require comprehensive evaluation. As discussed in Section 4.5, secondary metal leaching from bimetallic systems and the formation of toxic byproducts represent significant concerns that are rarely addressed in current studies [,]. Life cycle assessment studies are notably absent from the literature, leaving a critical gap in understanding the overall environmental footprint of these enhanced systems.

5.2. Future Research Directions

To propel the iron-persulfate systems from fundamental research toward practical implementation, future endeavors should pivot to addressing the core mechanistic and application bottlenecks through cutting-edge operando techniques, sustainable material design, and smart process integration. Especially, future research should prioritize the application of in situ/operando techniques (e.g., in situ Raman, X-ray absorption spectroscopy, Mössbauer spectroscopy) to dynamically monitor iron valence states, surface transformations, and reaction intermediates under operational conditions. This represents a crucial opportunity to resolve the mechanistic ambiguities highlighted in Section 3.1, Section 3.2, and Section 3.3, particularly for understanding interface electron transfer in heterostructures [,] and vacancy-mediated activation in defect-engineered catalysts [,]. Coupling these experimental approaches with computational methods (DFT, machine learning) can establish predictive structure-activity relationships and guide the rational design of next-generation catalysts with optimized iron cycling capabilities.
The current reliance on scarce or toxic metals presents a compelling opportunity to pioneer a new paradigm of catalysts designed for circularity and minimal environmental impact. Research should focus on developing earth-abundant, environmentally benign catalysts using green synthesis approaches. This includes transforming waste streams into valuable catalytic materials, thereby addressing both pollution control and waste management challenges. Waste-derived materials (e.g., sludge biochar [], spent battery materials [,]) offer promising avenues for reducing costs and environmental impacts while maintaining performance. Design strategies should emphasize structural robustness, easy separation, and regeneration capabilities to enhance practical applicability.
To realize this sustainable catalytic paradigm, several advanced material design strategies are emerging at the forefront. Beyond merely replacing critical metals, these approaches aim to maximize the activity and longevity of earth-abundant elements. For instance, engineering single-atom sites (e.g., Fe-N4) within conductive matrices can achieve near-optimal atom utilization while minimizing leaching, directly supporting the goal of minimal environmental impact [,]. Similarly, constructing confined nanoenvironments (e.g., within nanocages or MOFs) can protect active iron species from aggregation and poisoning, thereby enhancing the structural robustness and lifetime emphasized above [,]. Furthermore, the development of self-driven or energy-adaptive systems represents a pivotal step toward reducing operational energy costs, aligning economic and environmental sustainability [,]. Collectively, these advanced architectures provide the technological toolkit to build the next generation of high-performance yet sustainable iron-based persulfate catalysts.
The limitation of standalone AOPs in complex matrices can be transformed into an opportunity by developing intelligent, integrated process trains. Future work should explore the integration of iron-enhanced persulfate systems with complementary technologies (e.g., membrane filtration, biological treatment) to create synergistic treatment trains. Research on reactor design, process control, and energy optimization for continuous-flow operation is essential for technological maturation. Additionally, developing stimuli-responsive or adaptive systems that can maintain efficiency under fluctuating water quality conditions represents an important frontier.
The gap between laboratory benchmarks and field application underscores the critical need for standardized, environmentally focused evaluation protocols. Establishing standardized testing protocols for evaluating catalyst stability, metal leaching, and transformation product formation is crucial for meaningful cross-study comparisons. Research must expand beyond efficiency metrics to include thorough ecotoxicity assessments and life cycle analyses to ensure overall environmental benefits. Long-term field studies in real-world environments are necessary to validate laboratory findings and assess practical feasibility.
The significant energy footprint of external field-assisted strategies represents a major opportunity to align advanced oxidation processes with global carbon neutrality goals. Given the energy requirements of some enhancement strategies (e.g., electrochemical, UV assistance), future research should focus on coupling these systems with renewable energy sources. Solar-driven activation, particularly through photothermal and photovoltaic integration, represents a promising pathway toward carbon-neutral water treatment operations.

6. Conclusions

This review has systematically established that the kinetics of iron cycling constitute the central bottleneck in iron-persulfate oxidation systems. Through critical examination of three primary enhancement strategies—material design, system optimization, and external field assistance—we have established a comprehensive framework that links atomic-scale modifications to macroscopic performance improvements. Our analysis reveals that these strategies effectively enhance iron cycling through three convergent mechanisms: (1) facilitating electron transfer via conductive networks, redox mediators, or external energy fields; (2) suppressing iron precipitation through structural confinement, surface modification, and complexation; and (3) precisely regulating reaction pathway selection between radical and non-radical routes through electronic structure modulation. The integration of advanced characterization techniques and theoretical modeling has been crucial in elucidating these fundamental mechanisms at the molecular level. The remarkable progress in iron cycling enhancement has yielded substantial practical benefits, including unprecedented degradation efficiency for diverse contaminant classes, expanded operational pH windows, and significantly improved oxidant utilization rates. Specific material designs have demonstrated exceptional performance, with some heterostructured catalysts achieving up to 36-fold increases in degradation rates while maintaining stability over multiple cycles. However, our analysis reveals that future research must address several critical challenges to enable practical implementation. The development of standardized assessment protocols for evaluating long-term stability, environmental impacts, and economic feasibility is essential for meaningful technology advancement. Particular attention should be directed toward understanding the fate and transformation of catalyst components and reaction intermediates in complex environmental matrices, as well as managing sulfate accumulation in treated water. Looking forward, the field must embrace a holistic approach that balances performance optimization with sustainability considerations. The integration of renewable energy sources, waste-derived materials, and intelligent system designs represents a promising pathway toward sustainable water treatment applications. Future research should prioritize interdisciplinary collaborations that bridge materials science, environmental engineering, and computational methods to accelerate the development of next-generation iron-based persulfate systems. Ultimately, the rational design of iron-enhanced persulfate systems based on fundamental mechanistic understanding will enable transformative advances in water purification technologies. By addressing both scientific and practical challenges through integrated research efforts, we can realize the full potential of these systems for addressing emerging contaminant removal while maintaining environmental and economic sustainability.

Author Contributions

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

Funding

This research received no external funding.

Data Availability Statement

No new data were created or analyzed in this study. Data sharing is not applicable to this article.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Devendrapandi, G.; Liu, X.; Balu, R.; Ayyamperumal, R.; Arasu, M.V.; Lavanya, M.; Reddy, V.R.M.; Kim, W.K.; Karthika, P.C. Innovative remediation strategies for persistent organic pollutants in soil and water: A comprehensive review. Environ. Res. 2024, 249, 118404. [Google Scholar] [CrossRef] [PubMed]
  2. Qi, Y.; Cao, H.; Pan, W.; Wang, C.; Liang, Y. The role of dissolved organic matter during Per- and Polyfluorinated Substance (PFAS) adsorption, degradation, and plant uptake: A review. J. Hazard. Mater. 2022, 436, 129139. [Google Scholar] [CrossRef] [PubMed]
  3. 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]
  4. Cheng, M.; Zeng, G.; Huang, D.; Lai, C.; Xu, P.; Zhang, C.; Liu, Y. Hydroxyl radicals based advanced oxidation processes (AOPs) for remediation of soils contaminated with organic compounds: A review. Chem. Eng. J. 2016, 284, 582–598. [Google Scholar] [CrossRef]
  5. Kanakaraju, D.; Glass, B.D.; Oelgemoeller, M. Advanced oxidation process-mediated removal of pharmaceuticals from water: A review. J. Environ. Manag. 2018, 219, 189–207. [Google Scholar] [CrossRef]
  6. Poza-Nogueiras, V.; Rosales, E.; Pazos, M.; Angeles Sanroman, M. Current advances and trends in electro-Fenton process using heterogeneous catalysts—A review. Chemosphere 2018, 201, 399–416. [Google Scholar] [CrossRef]
  7. 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]
  8. 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]
  9. Matafonova, G.; Batoev, V. Recent advances in application of UV light-emitting diodes for degrading organic pollutants in water through advanced oxidation processes: A review. Water Res. 2018, 132, 177–189. [Google Scholar] [CrossRef]
  10. Zhao, G.; Zou, J.; Chen, X.; Liu, L.; Wang, Y.; Zhou, S.; Long, X.; Yu, J.; Jiao, F. Iron-based catalysts for persulfate-based advanced oxidation process: Microstructure, property and tailoring. Chem. Eng. J. 2021, 421, 127845. [Google Scholar] [CrossRef]
  11. Dong, H.; Xu, Q.; Lian, L.; Li, Y.; Wang, S.; Li, C.; Guan, X. Degradation of Organic Contaminants in the Fe(II)/Peroxymonosulfate Process under Acidic Conditions: The Overlooked Rapid Oxidation Stage. Environ. Sci. Technol. 2021, 55, 15390–15399. [Google Scholar] [CrossRef] [PubMed]
  12. Li, J.; Wan, Y.; Li, Y.; Yao, G.; Lai, B. Surface Fe(III)/Fe(II) cycle promoted the degradation of atrazine by peroxymonosulfate activation in the presence of hydroxylamine. Appl. Catal. B Environ. 2019, 256, 117782. [Google Scholar]
  13. Liu, Z.; Pan, S.; Xu, F.; Wang, Z.; Zhao, C.; Xu, X.; Gao, B.; Li, Q. Revealing the fundamental role of MoO2 in promoting efficient and stable activation of persulfate by iron carbon based catalysts: Efficient Fe2+/Fe3+cycling to generate reactive species. Water Res. 2022, 225, 119142. [Google Scholar] [CrossRef] [PubMed]
  14. Sheng, B.; Yang, F.; Wang, Y.; Wang, Z.; Li, Q.; Guo, Y.; Lou, X.; Liu, J. Pivotal roles of MoS2 in boosting catalytic degradation of aqueous organic pollutants by Fe(II)/PMS. Chem. Eng. J. 2019, 375, 121989. [Google Scholar] [CrossRef]
  15. Wang, Z.; Jiang, J.; Pang, S.; Zhou, Y.; Guan, C.; Gao, Y.; Li, J.; Yang, Y.; Qu, 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]
  16. Ahmed, H. Heterogeneous catalysts in advanced oxidation processes: A comprehensive review on antibiotic removal from wastewater. Sep. Purif. Technol. 2025, 374, 133670. [Google Scholar] [CrossRef]
  17. Hu, H.; Liu, J.; Zheng, X.; Zhao, K.; Lin, Y.; Xu, X.; Long, H.; Zhang, Y.; Wang, X.; Chen, D.; et al. Recent advances in iron-based catalyst-driven persulfate activation for organic pollutant degradation. J. Water Process Eng. 2025, 71, 107423. [Google Scholar] [CrossRef]
  18. Wang, H.; Guo, W.; Yin, R.; Du, J.; Wu, Q.; Luo, H.; Liu, B.; Sseguya, F.; Ren, N. Biochar-induced Fe(III) reduction for persulfate activation in sulfamethoxazole degradation: Insight into the electron transfer, radical oxidation and degradation pathways. Chem. Eng. J. 2019, 362, 561–569. [Google Scholar] [CrossRef]
  19. Xiong, Y.; Zhou, T.; Bao, J.; Du, J.; Faheem, M.; Luo, L. Degradation mechanism of Bisphenol S via hydrogen peroxide/persulfate activated by sulfidated nanoscale zero valent iron. Environ. Sci. Pollut. Res. 2023, 30, 83545–83557. [Google Scholar] [CrossRef]
  20. Gao, Y.; Duan, X.; Li, B.; Jia, Q.; Li, Y.; Fan, X.; Zhang, F.; Zhang, G.; Wang, S.; Peng, W. Fe containing template derived atomic Fe-N-C to boost Fenton-like reaction and charge migration analysis on highly active Fe-N4 sites. J. Mater. Chem. A 2021, 9, 14793–14805. [Google Scholar] [CrossRef]
  21. Xu, J.; Chen, J.; Zhong, Y.; Cao, L.; Zhang, X.; Wang, Z.; Chen, J.; Lin, S.; Xu, Q.; Chen, Y.; et al. Ultrathin MoS2 nanosheet-wrapped Fe3O4 nanocrystals synergistically activate peroxymonosulfate for enhanced removal of organic pollutants. Colloids Surf. A-Physicochem. Eng. Asp. 2023, 671, 131599. [Google Scholar] [CrossRef]
  22. Wang, X.; Wang, W.; Peng, J.; Wei, P.; Yan, G.; Cao, Z. Boosting effect of oxygen vacancy on peroxymonosulfate activation over CoWO4 for sulfapyridine removal. Sep. Purif. Technol. 2025, 371, 133285. [Google Scholar] [CrossRef]
  23. Meng, X.-R.; Chen, Q.; Wu, M.-F.; Wu, Q.; Wang, S.-Q.; Jin, L.-P.; Zhou, F.; Ma, R.-L.; Zou, J.-P. Nano-flowers FeS/MoS2 composites as a peroxymonosulfate activator for efficient p-chlorophenol degradation. Chin. J. Struct. Chem. 2025, 44, 100543. [Google Scholar] [CrossRef]
  24. Liang, J.; Guo, M.; Xue, Y.; Gu, J.-N.; Li, J.; Shi, F.; Guo, X.; Min, X.; Jia, J.; Li, K.; et al. Constructing magnetically separable manganese-based spinel ferrite from spent ternary lithium-ion batteries for efficient degradation of bisphenol A peroxymonosulfate activation. Chem. Eng. J. 2022, 435, 135000. [Google Scholar] [CrossRef]
  25. Liang, J.; Xue, Y.; Gu, J.-n.; Li, J.; Shi, F.; Guo, X.; Guo, M.; Min, X.; Li, K.; Sun, T.; et al. Sustainably recycling spent lithium-ion batteries to prepare magnetically separable cobalt ferrite for catalytic degradation of bisphenol A peroxymonosulfate activation. J. Hazard. Mater. 2022, 427, 127910. [Google Scholar] [CrossRef]
  26. Kang, H.; Lee, D.; Lee, K.; Kim, H.; Lee, H.; Kim, M.; 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]
  27. Luo, L.; Wang, Z.; Guo, Q.; Wei, X.; Hu, J.; Luo, Y.; Jiang, J. Evidence for the involvement of Fe(IV) in water treatment by Fe(III)-activated sulfite. Environ. Chem. Lett. 2022, 20, 91–99. [Google Scholar] [CrossRef]
  28. Wang, Z.; Qiu, W.; Pang, S.; Zhou, Y.; Gan, Y.; Guan, C.; Jiang, J. Further understanding the involvement of Fe(IV) in peroxydisulfate and peroxymonosulfate activation by Fe(II) for oxidative water treatment. Chem. Eng. J. 2019, 371, 842–847. [Google Scholar] [CrossRef]
  29. Li, S.; Tang, J.; Yu, C.; Liu, Q.; Wang, L. Efficient degradation of anthracene in soil by carbon-coated nZVI activated persulfate. J. Hazard. Mater. 2022, 431, 128581. [Google Scholar] [CrossRef]
  30. Yu, M.; Mao, X.; He, X.; Zheng, M.; Zhang, X.; Su, J.; Xi, B. Efficient degradation of sulfamethazine in a silicified microscale zero-valent iron activated persulfate process. Appl. Catal. B Environ. Energy 2022, 312, 121418. [Google Scholar] [CrossRef]
  31. He, D.; Cheng, Y.; Zeng, Y.; Luo, H.; Luo, K.; Li, J.; Pan, X.; Barcelo, D.; Crittenden, J.C. Synergistic activation of peroxymonosulfate and persulfate by ferrous ion and molybdenum disulfide for pollutant degradation: Theoretical and experimental studies. Chemosphere 2020, 240, 124979. [Google Scholar] [CrossRef]
  32. Yu, S.; Gu, X.; Lu, S.; Xue, Y.; Zhang, X.; Xu, M.; Qiu, Z.; Sui, Q. Degradation of phenanthrene in aqueous solution by a persulfate/percarbonate system activated with CA chelated-Fe(II). Chem. Eng. J. 2018, 333, 122–131. [Google Scholar] [CrossRef]
  33. Chen, Y.; Vu, H.C.; Miller, C.J.; Garg, S.; Pan, D.; Waite, T.D. Comparative Experimental and Computational Studies of Hydroxyl and Sulfate Radical-Mediated Degradation of Simple and Complex Organic Substrates. Environ. Sci. Technol. 2022, 56, 8819–8832. [Google Scholar] [CrossRef] [PubMed]
  34. Epold, I.; Trapido, M.; Dulova, N. Degradation of levofloxacin in aqueous solutions by Fenton, ferrous ion-activated persulfate and combined Fenton/persulfate systems. Chem. Eng. J. 2015, 279, 452–462. [Google Scholar] [CrossRef]
  35. Brienza, M.; Ahmed, M.; Escande, A.; Plantard, G.; Scrano, L.; Chiron, S.; Bufo, S.; Goetz, V. Relevance of a photo-Fenton like technology based on peroxymonosulphate for 17β-estradiol removal from wastewater. Chem. Eng. J. 2014, 257, 191–199. [Google Scholar] [CrossRef]
  36. Zhang, S.; Wei, Y.; Metz, J.; He, S.; Alvarez, P.; Long, M. Persistent free radicals in biochar enhance superoxide-mediated Fe(III)/Fe (II) cycling and the efficacy of CaO2 Fenton-like treatment. J. Hazard. Mater. 2022, 421, 126805. [Google Scholar] [CrossRef]
  37. Huang, Y.; Yu, X.; Gan, H.; Jiang, L.; Gong, H. Degradation and chlorination mechanism of fumaric acid based on SO4•−: An experimental and theoretical study. Environ. Sci. Pollut. Res. 2021, 28, 48471–48480. [Google Scholar] [CrossRef]
  38. Wang, S.; Wu, J.; Lu, X.; Xu, W.; Gong, Q.; Ding, J.; Dan, B.; Xie, P. Removal of acetaminophen in the Fe2+/persulfate system: Kinetic model and degradation pathways. Chem. Eng. J. 2019, 358, 1091–1100. [Google Scholar] [CrossRef]
  39. Xie, Y.; Guan, D.; Deng, Y.; Sato, Y.; Luo, Y.; Chen, G. Factors hindering the degradation of pharmaceuticals from human urine in an iron-activated persulfate system. J. Environ. Sci. 2024, 135, 130–148. [Google Scholar] [CrossRef]
  40. Xue, J.; Chen, M.; Wang, M. Potassium Hydrogen Phthalate Abatement by Activated Persulfate in ZVI-PS Systems. Water Air Soil Pollut. 2020, 231, 235. [Google Scholar] [CrossRef]
  41. Xue, W.; Li, J.; Chen, X.; Liu, H.; Wen, S.; Shi, X.; Guo, J.; Gao, Y.; Xu, J.; Xu, Y. Recent advances in sulfidized nanoscale zero-valent iron materials for environmental remediation and challenges. Environ. Sci. Pollut. Res. 2023, 30, 101933–101962. [Google Scholar] [CrossRef]
  42. Yi, Y.; Tu, G.; Tsang, P.; Fang, Z. Insight into the influence of pyrolysis temperature on Fenton-like catalytic performance of magnetic biochar. Chem. Eng. J. 2020, 380, 122518. [Google Scholar] [CrossRef]
  43. Zou, X.; Zhou, T.; Mao, J.; Wu, X. Synergistic degradation of antibiotic sulfadiazine in a heterogeneous ultrasound-enhanced Fe0/persulfate Fenton-like system. Chem. Eng. J. 2014, 257, 36–44. [Google Scholar] [CrossRef]
  44. Zou, J.; Ma, J.; Zhang, J. Comment on Electrolytic Manipulation of Persulfate Reactivity by Iron Electrodes for TCE Degradation in Groundwater. Environ. Sci. Technol. 2014, 48, 4630–4631. [Google Scholar] [CrossRef] [PubMed]
  45. Xu, G.; Liu, Q.; Mai, Z.; Wang, M.; Zhao, H.; Xu, K. Strategies for improving performance of iron-based catalysts in activating heterogeneous Fenton-like oxidation in pollutants degradation: From the perspective of materials structure design. Process Saf. Environ. Prot. 2024, 190, 794–820. [Google Scholar] [CrossRef]
  46. Wang, K.; Li, H.; Yu, Y.; Bai, Y.; Zhang, R. A mini review on construction strategies of novel metal-based catalysts with dual reaction centers (DRCs) employed in advanced oxidation processes (AOPs). J. Taiwan Inst. Chem. Eng. 2024, 164, 105678. [Google Scholar] [CrossRef]
  47. Wang, K.; Yang, Y.; Zhang, T.; Liang, Y.; Wang, Q. Degradation of methylene blue with magnetic Co-doped Fe3O4@FeOOH nanocomposites as heterogeneous catalysts of peroxymonosulfate. RSC Adv. 2019, 9, 17664–17673. [Google Scholar] [CrossRef]
  48. Li, T.; Pan, J.; Wang, X.; Fan, Z.; Shi, T.; Wang, L.; Gao, B. Insights into the fundamental role of Mo doping in facilitating the activation of peroxydisulfate by iron-based catalysts: Accelerating the generation of sulfate radicals. Chem. Eng. J. 2023, 477, 147000. [Google Scholar] [CrossRef]
  49. Liu, T.; Li, C.; Chen, X.; Chen, Y.; Cui, K.; Wang, D.; Wei, Q. Peroxymonosulfate Activation by Fe@N Co-Doped Biochar for the Degradation of Sulfamethoxazole: The Key Role of Pyrrolic N. Int. J. Mol. Sci. 2024, 25, 10528. [Google Scholar] [CrossRef]
  50. Duan, Y.; Liu, Y.; Wang, Y.; Wang, H.; Yin, W.; Xu, G. Recyclable Fe/S co-doped nanocarbon derived from metal-organic framework as a peroxymonosulfate activator for efficient removal of 2,4-dichlorophenol. Environ. Sci. Pollut. Res. 2023, 30, 6906–6918. [Google Scholar] [CrossRef]
  51. Jiang, L.; Liu, X.; Yang, M.; Jiang, R.; Gao, J.; Dai, Q. Catalytic degradation of bisphenol S via peroxymonosulfate activation by bimetallic Fe, Co-embedded N-enriched wheat straw biochar: Influencing factors, performance and mechanism. J. Water Process Eng. 2025, 74, 107864. [Google Scholar] [CrossRef]
  52. Li, X.; Wu, L.; Zhang, A.; Wu, S.; Lin, Y.; Yang, C. Cobalt doping amount determines dominant reactive species in peroxymonosulfate activation via porous carbon catalysts co-doped by cobalt and nitrogen. J. Environ. Sci. 2024, 138, 212–226. [Google Scholar] [CrossRef] [PubMed]
  53. Xiong, S.; Deng, Y.; Gong, D.; Tang, R.; Zheng, J.; Li, L.; Zhou, Z.; Su, L.; Liao, C.; Yang, L. Magnetically modified in-situ N-doped Enteromorpha prolifera derived biochar for peroxydisulfate activation: Electron transfer induced singlet oxygen non-radical pathway. Chemosphere 2021, 284, 131404. [Google Scholar] [CrossRef] [PubMed]
  54. Liu, T.; Wang, Q.; Li, C.; Cui, M.; Chen, Y.; Liu, R.; Cui, K.; Wu, K.; Nie, X.; Wang, S. Synthesizing and characterizing Fe3O4 embedded in N-doped carbon nanotubes-bridged biochar as a persulfate activator for sulfamethoxazole degradation. J. Clean. Prod. 2022, 353, 131669. [Google Scholar] [CrossRef]
  55. Ma, H.; Zhang, X.; Feng, G.; Ren, B.; Pan, Z.; Shi, Y.; Xu, R.; Wang, P.; Liu, Y.; Wang, G.; et al. Carbon nanotube membrane armed with confined iron for peroxymonosulfate activation towards efficient tetracycline removal. Sep. Purif. Technol. 2023, 312, 123319. [Google Scholar] [CrossRef]
  56. He, D.; Chen, P.; Zheng, P.; Yang, M.; Liu, L.; Yuan, S. Iron-nitrogen co-doped biochar (FeN@BC) as particle electrode for three-dimensional (3D) electro-peroxydisulfate process for tetracycline degradation. AIP Adv. 2024, 14, 035245. [Google Scholar] [CrossRef]
  57. Yang, B.; Tong, J.; Tong, J.; Peng, H.; Xiang, Y.; Ruan, M.; Chen, Z.; Yang, Z.; Xiong, W. Microwave synthesis of Fe-Cu diatomic active center MOF: Synergistic cyclic catalysis of persulfate for degrading norfloxacin. Environ. Sci.-Nano 2023, 10, 2778–2789. [Google Scholar] [CrossRef]
  58. Xiong, W.; Hu, F.; Liu, Y.; Nie, G.; Xiao, L. Core-shell FeMn@NG derived from cellulose supported Prussian blue analogs for peroxymonosulfate activation: Non-radical mechanism and ultra-low metal leaching. J. Environ. Chem. Eng. 2022, 10, 108523. [Google Scholar] [CrossRef]
  59. Lan, H.; Zhou, J.; Hou, Z.; An, X.; Liu, H.; Qu, J. Defect modulation of MOF-derived ZnFe2O4/CNTs microcages for persulfate activation: Enhanced nonradical catalytic oxidation. Chem. Eng. J. 2022, 431, 133369. [Google Scholar] [CrossRef]
  60. Shang, Y.; Chen, C.; Zhang, P.; Yue, Q.; Li, Y.; Gao, B.; Xu, X. Removal of sulfamethoxazole from water via activation of persulfate by Fe<sub>3</sub>C@NCNTs including mechanism of radical and nonradical process. Chem. Eng. J. 2019, 375, 122004. [Google Scholar]
  61. Li, Q.; Wang, L.; Zhang, L.; Xie, H. Rapid degradation of tetrabromobisphenol A under the UV/TiO2/KPS systems in alkaline aqueous solutions. Res. Chem. Intermed. 2019, 45, 757–768. [Google Scholar] [CrossRef]
  62. Zhong, C.; Wang, J.; Liu, Y.; Zhang, C.; Li, Z.; Liu, Z.; Zhao, C.; Yang, L.; Xu, X. Modulation of sulfur vacancies in MoS2/Fe3P enhancing non-radical pathway of the Fenton-like reaction for complex real water. Appl. Catal. B Environ. Energy 2025, 370, 125193. [Google Scholar] [CrossRef]
  63. He, J.; Yang, Y.; Hong, P.; Li, Y.; Xie, C.; Wu, Z.; Li, M.; Kong, L. Oxygen Vacancies of Mn/CeOx-H induced non-radical activation of peroxymonosulfate through electron mediation for bisphenol A degradation. J. Environ. Chem. Eng. 2023, 11, 111078. [Google Scholar] [CrossRef]
  64. Zhou, Z.; Tan, Y.; Yan, T.; Wu, X.; He, G.; Li, H.; Zhou, J.; Tong, H.; Yu, L.; Zeng, J. CoFe2O4-x/expanded graphite composite with abundant oxygen vacancies as a high-performance peroxymonosulfate activator for antibiotic degradation. Colloids Surf. A-Physicochem. Eng. Asp. 2025, 718, 136808. [Google Scholar] [CrossRef]
  65. Meng, B.; Fu, H.; Liu, S.-S.; Zhao, C.; Wang, F.; Li, S.; Guo, Y.; Meng, B.; Wang, J.-F.; Wang, C.-C. Polymerization deposition route governed by the coordination environment of interface asymmetric oxygen vacancy for efficient BPA removal via a Fenton-like reaction. Sep. Purif. Technol. 2025, 366, 132824. [Google Scholar] [CrossRef]
  66. Liang, X.; Wang, D.; Zhao, Z.; Li, T.; Gao, Y.; Hu, C. Coordination Number Dependent Catalytic Activity of Single-Atom Cobalt Catalysts for Fenton-Like Reaction. Adv. Funct. Mater. 2022, 32, 2203001. [Google Scholar] [CrossRef]
  67. Zhao, X.; Tong, J.; Bai, S.; Qian, J. Dual-metal-catalyzed Fenton-like reaction on CdxZn1-xS @biochar: Mechanistic insights into sulfide-metal interactions for water purification. Appl. Catal. B Environ. Energy 2025, 379, 125654. [Google Scholar] [CrossRef]
  68. Wang, Y.; Li, H.; Zong, Y.; Zhou, Z.; Ye, G.; Lei, Z.; Wu, D. Defect-Engineered Inert Interfaces on Iron-Rich Clay Minerals Boost Exclusive Electron Transfer Pathway in Fenton-Like Reactions. Adv. Funct. Mater. 2025, e12238. [Google Scholar] [CrossRef]
  69. Lai, L.; Zhou, H.; Hong, Y.; Luo, M.; Shi, Y.; Zhang, H.; Xiong, Z.; Yao, G.; Lai, B. Activation of peroxymonosulfate by FeVO3-x for the degradation of carbamazepine: Vanadium mediated electron shuttle and oxygen vacancy modulated interface chemistry. Chin. Chem. Lett. 2024, 35, 108580. [Google Scholar] [CrossRef]
  70. Fang, H.; Zhu, G.; Hu, Y.; Zhou, C.; Xu, M.; You, M.; Zhao, J.; Xue, S.; Zhang, J.; Liu, G. Flower-like C/CoO/Co3O4 heterostructures with abundant oxygen vacancies for effective peroxymonosulfate activation and tetracycline degradation. J. Solid State Chem. 2025, 350, 125469. [Google Scholar] [CrossRef]
  71. Ma, C.; Wang, J.; Wang, F.; Zhu, Y.; Li, Y.; Fan, X.; Zhang, F.; Zhang, G.; Peng, W. Facile synthesis of iron oxide supported on porous nitrogen doped carbon for catalytic oxidation. Sci. Total Environ. 2021, 785, 147296. [Google Scholar] [CrossRef]
  72. Sun, Y.; Ma, C.; Wu, D.; Liu, X.; Li, N.; Fan, X.; Li, Y.; Zhang, G.; Zhang, F.; Peng, W. Coating CoFe2O4 shell on Fe particles to increase the utilization efficiencies of Fe and peroxymonosulfate for low-cost Fenton-like reactions. Water Res. 2023, 244, 120542. [Google Scholar] [CrossRef]
  73. Wang, W.; Liu, Y.; Yue, Y.; Wang, H.; Cheng, G.; Gao, C.; Chen, C.; Ai, Y.; Chen, Z.; Wang, X. The Confined Interlayer Growth of Ultrathin Two-Dimensional Fe3O4 Nanosheets with Enriched Oxygen Vacancies for Peroxymonosulfate Activation. ACS Catal. 2021, 11, 11256–11265. [Google Scholar] [CrossRef]
  74. Song, Y.; Wang, H.; Shi, Z.; Cui, J.; Wu, D. Selective etching induced Fe3O4 cubes with orderly aligned porous structure for the effective lithium-ion storage and bisphenol A degradation. Sep. Purif. Technol. 2024, 334, 126054. [Google Scholar] [CrossRef]
  75. Ge, X.; Liu, J.; Song, X.; Wang, G.; Zhang, H.; Zhang, Y.; Zhao, H. Hierarchical iron containing γ-MnO2 hollow microspheres: A facile one-step synthesis and effective removal of As(III) via oxidation and adsorption. Chem. Eng. J. 2016, 301, 139–148. [Google Scholar] [CrossRef]
  76. Peng, T.; Zhang, H.; Xia, S.; Zhou, S.; Shi, Z.; Li, G.; Deng, L. MoS2 Nanosheets Anchored onto MIL-100(Fe)-Derived FeS2 as a Peroxymonosulfate Activator for Efficient Sulfamethoxazole Degradation: Insights into the Mechanism. ACS ES&T Water 2023, 3, 213–226. [Google Scholar]
  77. Wen, L.; Li, X.; Na, Y.; Chen, H.; Liu, M.; Yang, S.; Ding, D.; Wang, G.; Liu, Y.; Chen, Y.; et al. Surface reconstructed Fe@C1000 for enhanced Fenton-like catalysis: Sustainable ciprofloxacin degradation and toxicity reduction. Environ. Pollut. 2024, 345, 123534. [Google Scholar] [CrossRef]
  78. Lai, L.; Zhou, P.; Zhou, H.; Sun, M.; Yuan, Y.; Liu, Y.; Yao, G.; Lai, B. Heterogeneous Fe(III)/Fe(II) circulation in FeVO4 by coupling with dithionite towards long-lasting peroxymonosulfate activation: Pivotal role of vanadium as electron shuttles. Appl. Catal. B Environ. 2021, 297, 120470. [Google Scholar] [CrossRef]
  79. Li, Y.; Dong, H.; Xiao, J.; Li, L.; Hou, Y.; Chu, D.; Hou, X.; Xiang, S.; Dong, Q. Ascorbic acid-assisted iron silicate composite activated peroxydisulfate for enhanced degradation of aqueous contaminants: Accelerated Fe(III)/Fe(II) cycle and the interaction between iron and silicate. Chem. Eng. J. 2023, 455, 140733. [Google Scholar] [CrossRef]
  80. Yuan, Y.; Chen, S.; Yao, B.; Chen, A.; Pei, L.; Luo, S.; Zhou, Y. Fe3+-cysteine enhanced persulfate fenton-like process for quinclorac degradation: A wide pH tolerance and reaction mechanism. Environ. Res. 2023, 224, 115447. [Google Scholar] [CrossRef]
  81. Huang, M.; Zhu, C.; Zhu, F.; Fang, G.; Zhou, D. Mechanism of significant enhancement of VO2-Fenton-like reactions by oxalic acid for diethyl phthalate degradation. Sep. Purif. Technol. 2021, 279, 119671. [Google Scholar] [CrossRef]
  82. Li, Z.-Y.; Liu, Y.-L.; He, P.-N.; Zhang, X.; Wang, L.; Gu, H.-T.; Zhang, H.-C.; Ma, J. Further understanding the role of hydroxylamine in transformation of reactive species in Fe(II)/peroxydisulfate system. Chem. Eng. J. 2021, 418, 129464. [Google Scholar] [CrossRef]
  83. Li, X.; Ma, J.; Gao, Y.; Liu, X.; Wei, Y.; Liang, Z. Enhanced atrazine degradation in the Fe(III)/peroxymonosulfate system via accelerating Fe(II) regeneration by benzoquinone. Chem. Eng. J. 2022, 427, 131995. [Google Scholar] [CrossRef]
  84. Wu, J.; Cagnetta, G.; Wang, B.; Cui, Y.; Deng, S.; Wang, Y.; Huang, J.; Yu, G. Efficient degradation of carbamazepine by organo-montmorillonite supported nCoFe2O4-activated peroxymonosulfate process. Chem. Eng. J. 2019, 368, 824–836. [Google Scholar] [CrossRef]
  85. Ren, T.; Yang, S.; Jiang, Y.; Sun, X.; Zhang, Y. Enhancing surface corrosion of zero-valent aluminum (ZVAl) and electron transfer process for the degradation of trichloroethylene with the presence of persulfate. Chem. Eng. J. 2018, 348, 350–360. [Google Scholar] [CrossRef]
  86. Wan, Z.; Hu, X.; Li, C.; Zhang, J.; Wang, Q.; Fang, L.; Zhang, L.; Guo, Q.; Sun, D. Simultaneous oxidation and absorption of nitric oxide and sulfur dioxide by peroxymonosulfate activated by bimetallic metal-organic frameworks. J. Environ. Chem. Eng. 2023, 11, 109417. [Google Scholar] [CrossRef]
  87. Qiu, X.; Zhao, Y.; Jia, Z.; Li, C.; Jin, R.; Mutabazi, E. Fe and Zn co-doped carbon nanoparticles as peroxymonosulfate activator for efficient 2,4-dichorophenol degradation. Environ. Res. 2024, 240, 117313. [Google Scholar] [CrossRef]
  88. Wang, J.; Xie, T.; Liu, X.; Wu, D.; Li, Y.; Wang, Z.; Fan, X.; Zhang, F.; Peng, W. Enhanced redox cycle of Fe3+/Fe2+ on Fe@NC by boron: Fast electron transfer and long-term stability for Fenton-like reaction. J. Hazard. Mater. 2023, 445, 130605. [Google Scholar] [CrossRef]
  89. Xu, K.; Zhu, Z.; Hu, C.; Zheng, J.; Peng, H.; Liu, B. Superior degradation of organic contaminants by the UV/CoFe2O4/PI system: Kinetics, pathways, mechanisms and DFT calculation. Sep. Purif. Technol. 2024, 336, 126197. [Google Scholar] [CrossRef]
  90. Wang, X.; Li, J.; Chen, K.; Li, J.; Jia, Y.; Mei, Q.; Wang, Q. Facile synthesis of oxygen vacancies enriched ZnFe2O4 for effective photocatalytic peroxodisulfate activation. Sep. Purif. Technol. 2022, 303, 122205. [Google Scholar] [CrossRef]
  91. Chen, Y.; Zhao, D.; Sun, T.; Cai, C.; Dong, Y. The preparation of MoS2/8-FeOOH and degradation of RhB under visible light. J. Environ. Chem. Eng. 2023, 11, 110353. [Google Scholar] [CrossRef]
  92. He, Y.; Qian, J.; Wang, P.; Xie, T.; Dionysiou, D.D.; Lu, B.; Tang, S. Synergized selenium-vacancy heterogeneous interface and carbon nanotubes for insight into efficient oxidation of pollutants via photocatalytic peroxymonosulfate activation. Appl. Catal. B Environ. Energy 2023, 330, 122620. [Google Scholar] [CrossRef]
  93. Yuan, Y.; Wang, W.; Nie, M.; Yan, C.; Wang, P.; Ding, M. Visible light-mediated activation of periodate for bisphenol A degradation in the presence of Fe3+and gallic acid at neutral pH. Chem. Eng. J. 2024, 479, 147541. [Google Scholar] [CrossRef]
  94. An, L.; Kong, X.; Jiang, M.; Li, W.; Lv, Q.; Hou, X.; Liu, C.; Su, P.; Ma, J.; Yang, T. Photo-assisted natural chalcopyrite activated peracetic acid for efficient micropollutant degradation. Water Res. 2024, 257, 121699. [Google Scholar] [CrossRef] [PubMed]
  95. Hu, H.; Lin, H.; Chen, X.; Pan, Y.; Li, X.; Zhuang, Z.; Chen, H.; Wang, X.; Luo, M.; Zheng, K.; et al. Fe, Mn-Prussian blue analogue@MXene composites for efficient photocatalytic peroxydisulfate-activated degradation of tetracycline hydrochloride and photoelectrochemical desalination. Chem. Eng. J. 2023, 476, 146682. [Google Scholar] [CrossRef]
  96. Liu, L.; Zhan, R.; Zhang, M.; Li, J.; Wang, Z.; Mi, H.; Zhang, Y. Insights into the performance, mechanism, and ecotoxicity of levofloxacin degradation in CoFe2O4 catalytic peroxymonosulfate process. J. Environ. Chem. Eng. 2022, 10, 107435. [Google Scholar] [CrossRef]
  97. Wang, Y.; Ding, L.; Liu, C.; Lu, Y.; Wu, Q.; Wang, C.; Hu, Q. 0D/2D/2D ZnFe2O4/Bi2O2CO3/BiOBr double Z-scheme heterojunctions for the removal of tetracycline antibiotics by permonosulfate activation: Photocatalytic and non-photocatalytic mechanisms, radical and non-radical pathways. Sep. Purif. Technol. 2022, 283, 120164. [Google Scholar] [CrossRef]
  98. 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]
  99. Zhu, Y.; Wang, F.; Zhou, B.; Chen, H.; Yuan, R.; Zhang, Y.; Geng, H.; Liu, Y.; Wang, H. Photo-assisted Fe2+ modified molybdenum disulfide activated potassium persulfate to degrade sulfadiazine: Insights into the degradation pathway and mechanism from density functional theory. Chem. Eng. J. 2022, 435, 134904. [Google Scholar] [CrossRef]
  100. Shen, L.; Xu, Z.; Xiao, Z.; Zeng, Y.; Chen, K.; Lin, C.; Chen, W.; Chen, P.; Lv, W.; Liu, G. Insight into the degradation of difloxacin by heat/persulfate system: Mechanism, pathway and application. Sep. Purif. Technol. 2025, 362, 131661. [Google Scholar] [CrossRef]
  101. Yang, J.; Zeng, Z.; Huang, Z.; Cui, Y. Acceleration of Persulfate Activation by MIL-101(Fe) with Vacuum Thermal Activation: Effect of FeII/FeIII Mixed-Valence Center. Catalysts 2019, 9, 906. [Google Scholar] [CrossRef]
  102. Wang, B.W.; Gao, C.C.; Wang, W.S.; Zhao, H.B.; Zheng, Y.; Zheng, C.G. Effect of heating rate on chemical looping combustion of coal with Fe2O3 oxygen carrier. In Proceedings of the Asia-Pacific Power and Energy Engineering Conference (APPEEC), Wuhan, China, 25–28 March 2011. [Google Scholar]
  103. Chen, X.; Hu, X.; Gao, Y. Removal of NO in simulated flue with aqueous solution of peroxymonosulfate activated by high temperature and Fe(II). Chem. Eng. J. 2019, 359, 419–427. [Google Scholar] [CrossRef]
  104. Rodriguez-Chueca, J.; Giannakis, S.; Marjanovic, M.; Kohantorabi, M.; Gholami, M.R.; Grandjean, D.; de Alencastro, L.F.; Pulgarin, C. Solar-assisted bacterial disinfection and removal of contaminants of emerging concern by Fe2+-activated HSO5- vs. S2O82- in drinking water. Appl. Catal. B Environ. 2019, 248, 62–72. [Google Scholar] [CrossRef]
  105. Lin, H.; Zhong, X.; Ciotonea, C.; Fan, X.; Mao, X.; Li, Y.; Deng, B.; Zhang, H.; Royer, S. Efficient degradation of clofibric acid by electro-enhanced peroxydisulfate activation with Fe-Cu/SBA-15 catalyst. Appl. Catal. B Environ. 2018, 230, 1–10. [Google Scholar] [CrossRef]
  106. Lin, H.; Li, Y.; Mao, X.; Zhang, H. Electro-enhanced goethite activation of peroxydisulfate for the decolorization of Orange II at neutral pH: Efficiency, stability and mechanism. J. Taiwan Inst. Chem. Eng. 2016, 65, 390–398. [Google Scholar] [CrossRef]
  107. Wang, S.; Zhang, Y. Zero valent iron-electro-Fenton-peroxymonosulfate (ZVI-E-Fenton-PMS) process for industrial wastewater treatment. RSC Adv. 2023, 13, 15063–15076. [Google Scholar] [CrossRef]
  108. Chen, X.; Han, Y.; Gao, P.; Li, H. New insight into the mechanism of electro-assisted pyrite minerals activation of peroxymonosulfate: Synergistic effects, activation sites and electron transfer. Sep. Purif. Technol. 2021, 274, 118817. [Google Scholar] [CrossRef]
  109. Bai, Y.; Sun, X.; Dang, Y.; Yu, S.; Zhu, J.-J.; Zhou, Y. A self-circulating electro-fenton-like process over Fe3O4-CaO2 cathode for highly efficient degradation of levofloxacin. Chemosphere 2023, 313, 137520. [Google Scholar] [CrossRef]
  110. Zeng, H.; Zhao, X.; Zhao, F.; Park, Y.; Sillanpaa, M. Accelerated Fe3+/Fe2+ Cycle using Atomic H* on Pd/Al2O3: A Novel Mechanism for an Electrochemical System with Particle Electrode for Iron Sludge Reduction in the Fe2+/Peroxydisulfate Oxidation Process. Chem. Eng. J. 2020, 382, 122972. [Google Scholar] [CrossRef]
  111. Yan, S.; Zhang, X.; Zhang, H. Persulfate activation by Fe(III) with bioelectricity at acidic and near-neutral pH regimes: Homogeneous versus heterogeneous mechanism. J. Hazard. Mater. 2019, 374, 92–100. [Google Scholar] [CrossRef]
  112. Jiang, X.; Huang, Y.; Xu, W.; Liu, Z.; Song, L.; Ruan, B.; Lin, Z.; Zhang, H.; Wang, H.; Yeung, K.L. Synergistic enhancement of sulfamethoxazole degradation via nitrogen-sulfur co-doped iron-based biochar in a 3D electro-Fenton system. Sep. Purif. Technol. 2025, 375, 133881. [Google Scholar] [CrossRef]
  113. Long, X.; Huang, R.; Li, Y.; Wang, J.; Zhang, M.; Zhang, I.Y. Understanding the electro-cocatalytic peroxymonosulfate-based systems with BDD versus DSA anodes: Radical versus nonradical dominated degradation mechanisms. Sep. Purif. Technol. 2023, 309, 123120. [Google Scholar] [CrossRef]
  114. Pang, Y.; Ruan, Y.; Feng, Y.; Diao, Z.; Shih, K.; Hou, L.A.; Chen, D.; Kong, L. Ultrasound assisted zero valent iron corrosion for peroxymonosulfate activation for Rhodamine-B degradation. Chemosphere 2019, 228, 412–417. [Google Scholar] [CrossRef]
  115. Bao, X.; Zhang, Y.; Zhang, M.; Luo, G.; Deng, Y.; He, Y.; Zhang, Y. Ultrasound irritation enhanced activation of peroxymonosulfate with Fe0 for humic acid removal: Less formation of Fe-HA complex. Desalin. Water Treat. 2021, 237, 108–120. [Google Scholar] [CrossRef]
  116. Chi, N.; Liu, J.; Feng, L.; Guo, Z.; Chen, Y.; Pan, T.; Zheng, H. FeS redox power motor for PDS continuous generation of active radicals on efficient degradation and removal of diclofenac: Role of ultrasonic. Chemosphere 2022, 300, 134574. [Google Scholar] [CrossRef] [PubMed]
  117. Xiang, W.; Chen, H.; Zhong, Z.; Zhang, C.; Lu, X.; Huang, M.; Zhou, T.; Yu, P.; Zhang, B. Efficient degradation of carbamazepine in a neutral sonochemical FeS/persulfate system based on the enhanced heterogeneous-homogeneous sulfur-iron cycle. Sep. Purif. Technol. 2022, 282, 120041. [Google Scholar] [CrossRef]
  118. Lv, Z.; You, H.; Leng, H.; Li, W.; Yang, S.; Li, Z.; Zhu, J.; Zhang, G. Revealing the catalytic behavior of different types ZVI (powder Fe0, sponge Fe1, foam Fe2) in US/Fex/PDS and its effect on the dewaterability of waste activated sludge. Chem. Eng. J. 2024, 500, 157010. [Google Scholar] [CrossRef]
  119. Liu, J.; Zhou, J.; Ding, Z.; Zhao, Z.; Xu, X.; Fang, Z. Ultrasound irritation enhanced heterogeneous activation of peroxymonosulfate with Fe3O4 for degradation of azo dye. Ultrason. Sonochem. 2017, 34, 953–959. [Google Scholar] [CrossRef]
  120. Wang, J.; Liang, M. Fe3O4@BPC nanocatalyst: Enhanced activation of peroxymonosulfate for effective degradation of norfloxacin. Mater. Lett. 2023, 352, 135223. [Google Scholar] [CrossRef]
  121. Liu, C.; Zhang, J.; Wang, Q.; Zhang, G. Positive Effects of Ultrasound-Assisted Catalytic Activation of Peroxymonosulfate Combined with Fe-Niox Catalyst for the Orangeii Degradation In Acidic Solution. Fresenius Environ. Bull. 2020, 29, 7002–7013. [Google Scholar]
  122. Bazipour, F.; Jorfi, S.; Maleki, H.; Babaei, A. Catalytic Degradation of Acid Orange 7 Using CoFe2O4@Biochar Heterogeneous Catalytic Ozonation Process in Aqueous Solutions. Water Air Soil Pollut. 2024, 235, 811. [Google Scholar] [CrossRef]
  123. Hong, Q.; Liu, C.; Wang, Z.; Li, R.; Liang, X.; Wang, Y.; Zhang, Y.; Song, Z.; Xiao, Z.; Cui, T.; et al. Electron transfer enhancing Fe(II)/Fe(III) cycle by sulfur and biochar in magnetic FeS@biochar to active peroxymonosulfate for 2,4-dichlorophenoxyacetic acid degradation. Chem. Eng. J. 2021, 417, 129238. [Google Scholar] [CrossRef]
  124. Lee, Y.; Lo, S.; Chiueh, P.; Liou, Y.; Chen, M. Microwave-hydrothermal decomposition of perfluorooctanoic acid in water by iron-activated persulfate oxidation. Water Res. 2010, 44, 886–892. [Google Scholar] [CrossRef] [PubMed]
  125. Sühnholz, S.; Gawel, A.; Kopinke, F.; Mackenzie, K. Evidence of heterogeneous degradation of PFOA by activated persulfate-FeS as adsorber and activator. Chem. Eng. J. 2021, 423, 130102. [Google Scholar] [CrossRef]
  126. Tran, T.; Abrell, L.; Brusseau, M.; Chorover, J. Iron-activated persulfate oxidation degrades aqueous Perfluorooctanoic acid (PFOA) at ambient temperature. Chemosphere 2021, 281, 130824. [Google Scholar] [CrossRef]
  127. Zhang, Y.; Sun, J.; Guo, Z.; Zheng, X.; Guo, P.; Xu, J.; Lei, Y. The decomplexation of Cu-EDTA by electro-assisted heterogeneous activation of persulfate via acceleration of Fe(II)/Fe(III) redox cycle on Fe-MOF catalyst. Chem. Eng. J. 2022, 430, 133025. [Google Scholar] [CrossRef]
  128. Zhao, J.; Zhong, P.; Luo, W.; Zhang, S.; Xu, S.; Yu, Q.; Qiu, X. Insight into in-situ chemical oxidation by Fe(II)-containing minerals: The role of inherent Fe(II)-OH in Fe(II)-Al LDHs. Chem. Eng. J. 2022, 433, 133835. [Google Scholar] [CrossRef]
  129. Zhao, Y.; Huang, B.; An, H.; Dong, G.; Feng, J.; Wei, T.; Ren, Y.; Ma, J. Enhanced activation of peroxymonosulfate by Sr-doped LaFeO3 perovskite for Orange I degradation in the water. Sep. Purif. Technol. 2021, 256, 117838. [Google Scholar] [CrossRef]
  130. Wang, Y.; Chen, S.-Y.; Yang, X.; Huang, X.-F.; Yang, Y.-H.; He, E.-K.; Wang, S.; Qiu, R.-L. Degradation of 2,2′,4,4′-tetrabromodiphenyl ether (BDE-47) by a nano zerovalent iron-activated persulfate process: The effect of metal ions. Chem. Eng. J. 2017, 317, 613–622. [Google Scholar] [CrossRef]
  131. Liu, B.; Song, W.; Wu, H.; Xu, Y.; Sun, Y.; Yu, Y.; Zheng, H.; Wan, S. Enhanced oxidative degradation of norfloxacin using peroxymonosulfate activated by oily sludge carbon-based nanoparticles CoFe2O4/OSC. Chem. Eng. J. 2020, 400, 125947. [Google Scholar] [CrossRef]
  132. Li, Y.; Zhao, X.; Yan, Y.; Yan, J.; Pan, Y.; Zhang, Y.; Lai, B. Enhanced sulfamethoxazole degradation by peroxymonosulfate activation with sulfide-modified microscale zero-valent iron (S-mFe0): Performance, mechanisms, and the role of sulfur species. Chem. Eng. J. 2019, 376, 121302. [Google Scholar] [CrossRef]
  133. Pan, T.; Wang, Y.; Yang, X.; Huang, X.-F.; Qiu, R.-L. Gallic acid accelerated BDE47 degradation in PMS/Fe(III) system: Oxidation intermediates autocatalyzed redox cycling of iron. Chem. Eng. J. 2020, 384, 123248. [Google Scholar] [CrossRef]
  134. Nie, G.; Huang, J.; Hu, Y.; Ding, Y.; Han, X.; Tang, H. Heterogeneous catalytic activation of peroxymonosulfate for efficient degradation of organic pollutants by magnetic Cu/Fe3O4 submicron composites. Chin. J. Catal. 2017, 38, 227–239. [Google Scholar] [CrossRef]
  135. Feng, Y.; Lee, P.-H.; Wu, D.; Shih, K. Surface-bound sulfate radical-dominated degradation of 1,4-dioxane by alumina-supported palladium (Pd/Al2O3) catalyzed peroxymonosulfate. Water Res. 2017, 120, 12–21. [Google Scholar] [CrossRef]
  136. Lee, H.; Kim, H.-I.; Weon, S.; Choi, W.; Hwang, Y.S.; Seo, J.; Lee, C.; Kim, J.-H. Activation of Persulfates by Graphitized Nanodiamonds for Removal of Organic Compounds. Environ. Sci. Technol. 2016, 50, 10134–10142. [Google Scholar] [CrossRef]
  137. Zhuo, S.-N.; Ren, H.-Y.; Cao, G.-L.; Xie, G.-J.; Xing, D.-F.; Ren, N.-Q.; Liu, B.-F. Highly efficient activation of persulfate by encapsulated nano-Fe0 biochar for acetaminophen degradation: Rich electron environment and dominant effect of superoxide radical. Chem. Eng. J. 2022, 440, 135947. [Google Scholar] [CrossRef]
  138. Du, J.; Guo, W.; Che, D.; Ren, N. Weak magnetic field for enhanced oxidation of sulfamethoxazole by Fe0/H2O2 and Fe0/persulfate: Performance, mechanisms, and degradation pathways. Chem. Eng. J. 2018, 351, 532–539. [Google Scholar] [CrossRef]
  139. Li, Z.; Wang, F.; Zhang, Y.; Lai, Y.; Fang, Q.; Duan, Y. Activation of peroxymonosulfate by CuFe2O4-CoFe2O4 composite catalyst for efficient bisphenol a degradation: Synthesis, catalytic mechanism and products toxicity assessment. Chem. Eng. J. 2021, 423, 130093. [Google Scholar] [CrossRef]
  140. Pu, M.; Ma, Y.; Wan, J.; Wang, Y.; Wang, J.; Brusseau, M.L. Activation performance and mechanism of a novel heterogeneous persulfate catalyst: Metal-organic frameworkMIL-53(Fe) with FeII/FeIII mixed-valence coordinatively unsaturated iron center. Catal. Sci. Technol. 2017, 7, 1129–1140. [Google Scholar] [CrossRef]
  141. Lu, J.; Zhou, Y.; Lei, J.; Ao, Z.; Zhou, Y. Fe3O4/graphene aerogels: A stable and efficient persulfate activator for the rapid degradation of malachite green. Chemosphere 2020, 251, 126402. [Google Scholar] [CrossRef]
  142. Zhao, Y.; Song, M.; Cao, Q.; Sun, P.; Chen, Y.; Meng, F. The superoxide radicals’ production persulfate activated with CuFe2O4@ Biochar composites to promote the redox pairs cycling for efficient degradation of o-nitrochlorobenzene in soil. J. Hazard. Mater. 2020, 400, 122887. [Google Scholar] [CrossRef]
  143. Qu, S.; Li, C.; Sun, X.; Wang, J.; Luo, H.; Wang, S.; Ta, J.; Li, D. Enhancement of peroxymonosulfate activation and utilization efficiency via iron oxychloride nanosheets in visible light. Sep. Purif. Technol. 2019, 224, 132–141. [Google Scholar] [CrossRef]
  144. Wang, S.; Xu, W.; Wu, J.; Gong, Q.; Xie, P. Improved sulfamethoxazole degradation by the addition of MoS2 into the Fe2+/peroxymonosulfate process. Sep. Purif. Technol. 2020, 235, 116170. [Google Scholar] [CrossRef]
  145. Luo, H.; Zhou, X.; Guo, X.; Fang, Z.; Chen, Q.; Zhou, J. WS2 as highly active co-catalyst for the regeneration of Fe(II) in the advanced oxidation processes. Chemosphere 2021, 262, 128067. [Google Scholar] [CrossRef] [PubMed]
  146. Li, Y.; Feng, Y.; Yang, B.; Yang, Z.; Shih, K. Activation of peroxymonosulfate by molybdenum disulfide-mediated traces of Fe(III) for sulfadiazine degradation. Chemosphere 2021, 283, 131212. [Google Scholar] [CrossRef] [PubMed]
  147. Zhang, Q.; Sun, X.; Dang, Y.; Zhu, J.-J.; Zhao, Y.; Xu, X.; Zhou, Y. A novel electrochemically enhanced homogeneous PMS-heterogeneous CoFe2O4 synergistic catalysis for the efficient removal of levofloxacin. J. Hazard. Mater. 2022, 424, 127651. [Google Scholar] [CrossRef]
  148. Jiang, F.; Li, Y.; Zhou, W.; Yang, Z.; Ning, Y.; Liu, D.; Tang, Z.; Yang, S.; Huang, H.; Wang, G. Enhanced degradation of monochlorobenzene in groundwater by ferrous iron/persulfate process with cysteine. Chem. Eng. J. 2020, 387, 124048. [Google Scholar] [CrossRef]
  149. Wang, C.; Liu, H.; Li, X.; Cao, Y.; Jia, K. Degradation of organic pollutants in natural magnetite-activated peroxymonosulfate system enhanced by hydroxylamine: Catalytic performance, reaction pathway, and enhancement mechanism. J. Water Process Eng. 2024, 66, 106000. [Google Scholar] [CrossRef]
  150. Li, R.; Lu, X.; Yan, B.; Li, N.; Chen, G.; Cheng, Z.; Hou, L.A.; Wang, S.; Duan, X. Sludge-derived biochar toward sustainable Peroxymonosulfate Activation: Regulation of active sites and synergistic production of reaction oxygen species. Chem. Eng. J. 2022, 440, 135897. [Google Scholar] [CrossRef]
  151. Tan, C.; Cheng, X.; Xu, T.; Chen, K.; Xiang, H.; Su, L. Crystalline boron significantly accelerates Fe(III)/PMS reaction as an electron donor: Iron recycling, reactive species generation, and acute toxicity evaluation. Chem. Eng. J. 2023, 452, 139154. [Google Scholar] [CrossRef]
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.

Article Metrics

Citations

Article Access Statistics

Multiple requests from the same IP address are counted as one view.