Next Article in Journal
Conversion of Materials and Energy in Anaerobic Digestion of Sewage Sludge with High-Pressure Homogenization Pretreatment
Next Article in Special Issue
Electrocatalytic Ni-Co Metal Organic Framework for Efficient Urea Oxidation Reaction
Previous Article in Journal
Sedimentary Facies Types and Their Control of Reservoirs in the Lower Jurassic Lacustrine Facies Shale of the Lianggaoshan Formation, Northeastern Sichuan Basin, China
Previous Article in Special Issue
Progress in Electroreduction of CO2 to Form Various Fuels Based on Zn Catalysts
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Fenton: A Systematic Review of Its Application in Wastewater Treatment

by
Matheus Pimentel Prates
1,
Suzana Maria de Oliveira Loures Marcionílio
1,*,
Karine Borges Machado
2,
Danyelle Medeiros de Araújo
3,
Carlos A. Martínez-Huitle
3,4,*,
Arizeu Luiz Leão Arantes
1 and
José Eduardo Ferreira da Silva Gadêlha
1
1
Instituto Federal Goiano—Departamento de Agroquímica, Campus Rio Verde, Rio Verde 75901-970, GO, Brazil
2
Unidade Universitária de Anápolis (UnuCET), State University of Goiás, Anápolis 75132-903, GO, Brazil
3
National Institute for Alternative Technologies of Detection, Toxicological Evaluation and Removal of Micropollutants and Radioactives (INCT-DATREM), Institute of Chemistry, UNESP, Araraquara 14800-900, SP, Brazil
4
Renewable Energies and Environmental Sustainability Research Group, Federal University of Rio Grande do Norte, Campus Universitário, Av. Salgado Filho 3000, Lagoa Nova, Natal 59078-970, RN, Brazil
*
Authors to whom correspondence should be addressed.
Processes 2023, 11(8), 2466; https://doi.org/10.3390/pr11082466
Submission received: 6 January 2023 / Revised: 8 August 2023 / Accepted: 10 August 2023 / Published: 16 August 2023
(This article belongs to the Special Issue State of the Art of Waste Utilization and Resource Recovery)

Abstract

:
The use of new technologies for the removal of pollutants from wastewater has become globally necessary due to the complexity and facilities defined by conventional treatments. Advanced oxidative processes, specifically the Fenton process, have become widely applied given their low cost and ease of use. Therefore, this study aimed to evaluate the progression of the scientific publications on the implementation of Fenton process, investigating their space–time evolution. Additionally, useful solutions, trends, and gaps in the applications for the removal of pollutants with this methodology were identified, and also different remediation strategies and the design of new treatments for wastewaters were identified within this scientometric analysis. Bibliometric research was conducted in two scientific databases, Web of Science and Scopus, from 2011 to 2022, and we identified 932 and 1263 studies with the word “Fenton,” respectively. When these publications are associated with the treatment of alternative effluents, an increase in publications from 2011 (r = 0.95, p < 0.001) and 2013 (r = 0.93, p < 0.001) was observed when analyzing both databases, indicating the relevance of the theme. Among these studies, several of them were conducted on the bench scale (89.8% and 98.3%, Web of Science and Scopus, respectively) and in aqueous matrix (97.8% and 98.4%, Web of Science and Scopus, respectively), with being China the main country with publications associated with these words (28.33% and 41.9%), while Brazil is related to 3.65% and 2.29% of the total studies in Web of Science and Scopus, respectively. In addition, this review provides a guideline for new applications for different species in the matrices and describes the evolution of technological solutions to meet Sustainable Development Goal 6: clean water and sanitation.

1. Introduction

According to the 2017 UNESCO report on water-resource development [1] contemporary production and lifestyles are closely associated with strict demands for high drinking water quality, but production and lifestyle have led to serious water pollution. Therefore, the same report states that about 80% of the wastewater is still globally returned to the environment without adequate treatment. Then, to mitigate the resulting pollution, regulatory agencies currently require on-site treatment of generated effluents. Several technologies are used for effluent treatment, including filtration, reverse osmosis, adsorption, and coagulation–flocculation [2,3,4,5]. However, new processes are being established to compare the results and increase the efficiency of the removal of persistent, refractory, and nonbiodegradable pollutants, such as drugs, pesticides, dyes, organic solvents, phenols, oily effluents, and agroindustrial wastes, such as stillage [6,7].
To degrade certain types of pollutants, chemical treatment, which generates oxidizing agents, particularly hydroxyl radicals (OH), must be used. Currently, the most effective methods for degrading organic pollutants in wastewater and contaminated soil are advanced oxidative processes (AOPs), which include the Fenton approach. The Fenton process produces OH through a homogeneous reaction (takes place via the reaction from a mixture of hydrogen peroxide and ferrous salts). These oxidants are highly reactive and nonselective, reacting with a wide range of organic compounds and, consequently, degrading them into water, carbon dioxide, and inorganic ions [8]. For the by-products, the type of reaction that occurs between OH and the intermediate products will depend on the structure of the target compound.
Treatment processes based on the Fenton reaction have been effectively applied to treat various types of industrial effluents and municipal sewage treatment plants, as a single-, pre- and/or post-treatment methods [9]. For example, it can be used to treat leachate from landfill, from vinasse, as post-treatment of biodigested vinasse, in the treatment of effluents from the wood industry, in the removal of endocrine disruptors as well as to remove other environmental micro pollutants from the pharmaceutical industry, from hospital wastes, from university laboratories, from pesticides residues, and others [9,10].
The systematic literature review has become an established approach for comprehending the current state of knowledge on a particular topic [11,12,13]. This methodology enables not only the summary of the existing literature but also an assessment of spatial and temporal trends, as well as the identification of key biases and gaps, in the published research. It is important to note that some reviews do not bring details about the search methodologies to compose the review articles, which differs from the methodology adopted in the present work. For example, a general review on about the “Recent progress in Fenton/Fenton-like reactions for the removal of antibiotics in aqueous environments” was published in 2022 by Jiang and co-authors [14]. Nevertheless, the article outlines in general terms the applications of various Fenton oxidation technologies for the removal of antibiotics, the advantages, and disadvantages of these Fenton reactions in antibiotic removal. On the other hand, it described that Fenton-like oxidation requires less severe reaction conditions than Fenton oxidation, and it was more applicable to a wider range of processes, exhibiting lower economic costs and much higher cycle times. Additionally, it was emphasized that there is a simple and easy recycling method for the catalyst. Meanwhile, Wang and Zhuan [15] have also published the review of AOPs for the degradation of antibiotics without a specific bibliometric methodology. Conversely, the recent review follow the bibliometric methodology reported by other authors [16,17]. With this in mind, in this work a systematic literature review of the Fenton approach in wastewater treatment was carried out, with the aim of outlining its global use, reaction details, forms of data analysis, and temporal variations.

2. Methods

To identify relevant articles, a systematic literature review was conducted using the Web of Science (WoS, accessed on 3 January 2023) and SCOPUS databases (www.scopus.com, accessed on 23 January 2023). The search, in both databases, initially focused on papers with titles containing the terms Fenton* AND “industrial effluent*” OR wastewater* OR “water residual*” OR “persistent* compound*” OR “recalcitrant* compound*” OR “organic* load* “ OR “chemistry oxygen* demand*” OR “vinasse” OR “hexaferrite nanoparticle*” OR “Fe2O3 nanoparticle *” until the year 2022. Only scientific articles were considered. Scopus is a comprehensive international database with unique advanced tools, such as ranking of journals, author profiles, number of articles published by a journal in a given year, and frequency of use of scientific terms.
The year of publication, journal, and country of the first author for each article containing the above terms were identified. The title, abstract, and keywords of the articles focusing on Fenton treatments were reviewed to obtain the following information: (i) the study scale (bench, pilot, or industrial); (ii) the use of Fenton alone or in combination with other treatments; (iii) the type of matrix (aqueous, soil, or air); (iv) reaction conditions (traditional, Fenton-like, nanoparticles, and various iron sources); (v) chemical oxygen demand (COD) and total organic carbon (TOC) removal percentages; (vi) compounds degraded by Fenton (dyes, pharmaceuticals, hygiene and cleaning products, phenol, pesticides, organic and inorganic compounds, vinasse, and other specific compounds); and (vii) the origin of the degraded effluents (industrial, hospital, sewage treatment plant, teaching or research institution, synthetic solution, or other specific origins).
To assess the change in the number of publications on Fenton over time, the Pearson correlation between the number of articles and the years that were published was determined. The article count was standardized by dividing the number of articles on Fenton published, each year, by the total number of articles in the WoS and Scopus databases for that year. We also used Pearson correlations to examine the trends over time in the types of effluents and compounds that Fenton has been able to degrade.

3. Results and Discussions

3.1. Time Trends in Studies Using Fenton

Considering WoS and Scopus databases, the search identified a total of 1073 and 1653 articles in indexed journals, respectively, until 2022, with the earliest indexed publication on the theme in Scopus dating back to 1992. Among these articles, 141 and 390 contained the term “Fenton” in their title but these did not employ the technique in their study. These papers were excluded from further analysis as they were not aligned with the objectives of this systematic review. In total, 932 and 1263 articles (WoS and Scopus databases, respectively) on the implementation of the Fenton technique to remove any form of pollutant were considered.
In regard to the use of the WoS database, it is important to indicate that WoS is mainly designed for researchers to find published literature, and it could be restricted for bibliometric studies. Then, it is not suitable to directly use all different types and levels of the WoS database for bibliometric study. For example, Data Citation Index, Derwent Innovations Index, Zoological Record, Social Sciences Citation Index, Arts & Humanities Citation Index, Conference Proceedings Citation Index—Social Sciences & Humanities, Book Citation Index—Social Sciences & Humanities, Current Chemical Reactions, and Index Chemicus. In fact, a minor number of articles was identified by the WoS search. Then, the information was collected once again with the same methodology previously used but the Scopus database was utilized.
Comparing the results obtained in the two databases (Figure 1), it is observed that WoS is less comprehensive than Scopus. This may be due to the Scopus database having a larger, internationally diverse database as well as a range of intelligent tools that can be used to aid the research process, such as journal rankings, author profiles, the number of articles published by a journal in a given year, and the frequency of use of scientific terms. Therefore, Scopus was used to elaborate the systematic review, examining the results obtained. Additionally, WoS results were reported as Supplementary Materials (SM) in order to illustrate the similarities.
Figure 1 illustrates the increase in the number of articles over time (Pearson’s r = 0.95 and 0.93, for WoS and Scopus databases, respectively, with p < 0.001); a more pronounced growth in publications is attained since 2013 and, notably, in the years 2020 and 2022.
The findings revealed a trend over a period of thirty years in which the number of publications featuring AOPs has increased. This pattern is consistent with the macro-level explorations of chemical, electrochemical, photochemical, and photoelectrochemical technologies designed to remove organic pollutants from wastewaters [18]. The growing demand for the treatment of effluents that contain complex and recalcitrant pollutants is driving the use of the Fenton method [19]. Furthermore, the Fenton process is a versatile technique that can be combined with other technologies, as highlighted by Ramos et al. (2021) [20], who described the use of Fenton in a systematic review on textile dye treatment. Fenton is a homogeneous reaction, which occurs via a chemical reaction of the mixture of Fe2+ and H2O2, which leads to the rise of homogeneous hydroxyl radicals (OH) [9]. The use of Fenton is due to the growing demand for wastewater treatment containing increasingly complex and difficult to remove recalcitrant pollutants. It means that, with this type of oxidant, it is possible to treat different and varied ranges of residues, allowing for a low cost and reaction time, as well as avoiding the use of large technological devices [9].
From 2017 to 2022, researchers have incorporated statistical tools such as Takeuchi’s method using the Mini Tap program, as well as the response surface methodology (RSM) with designs including the Box–Behnken design, Box–Wilson statistical design, central composite design (CCD), Central Composite Design in MINITAB®16 software, and other arrangements, such as the orthogonal array test and single-factor experiments, orthogonal experiments, and the regression quadratic equation [21,22,23,24,25,26,27,28]. Furthermore, the use of new reaction forms has increased, as described below.
Table 1 displays the nine categories of pollutants that were degraded using Fenton during 1992–2022, across 1263 studies. Among these studies, 90 focused on the use of Fenton to concurrently degrade various compounds, specifically those falling under the categories of pharmaceuticals, hygiene/cleaning products, and dyes.
The table above displays the results of studies that focused on the removal of pollutants using the Fenton technique. Among these studies, the majority focused on the removal of dyes (227; 18%), organic compounds (216; 17%), and pharmaceuticals (114; 9%); fewer were focused on removing phenol (62; 5%), inorganic compounds/metals (40; 3%), pesticides (22; 1.7%), hygiene and cleaning products (11; 0.87%), and vinasse (5; 0.39%). However, a large proportion of the studies (418; 33.1%) examined the use of the Fenton technique to remove a diverse range of other compounds, including solvents, polymers, organic molecules, wastewater from kitchens, dairies, slaughterhouses, tanneries, petroleum industry, mining, biodiesel, edible oil refineries, tertiary pulp treatment, wood industries and municipal effluents, bacteria, oilseed, and olive-oil industries [29,30]. We noted an increase in the number of articles over time for all compounds evaluated (Table 1); however, this increase occurred to a greater extent for dyes (r = 0.85; p < 0.001), organic compounds (0.87, p < 0.001), pharmaceuticals (r = 0.81; p < 0.001), phenol (r = 0.81; p < 0.001), inorganic compounds/metals (r = 0.82, p < 0.001), pesticides (r =0.59; p < 0.001), hygiene and cleaning products (r = 0.42; p = 0.02), vinasse (r = 0.32; p = 0.09), and other types of specific composite products (r = 0.69; p < 0.001). These results are in agreement with the bibliometric analysis of WoS (see Figures S1 and S2 in the Supplementary Materials)
Regarding the treatment of wastewater with dyes, studies on this topic have increased in the last 10 years of the study period (2012–2022; r = 0.85 with time) [31,32,33]. Studies on the removal of organic compounds increased after 2015 (r = 0.87), showing that organic compounds were the most commonly used matrix in Fenton studies [34,35,36] in comparison with pharmaceuticals [37,38,39], phenols [40,41], pesticides [42,43], inorganic compounds [44,45], hygiene and cleaning products [46,47].
Figure 2 presents the sources of the pollutants listed in Table 1. However, 132 studies did not specify the origin of the treated matrix. Among those that were identified, 16 studies used the Fenton method to simultaneously degrade products from different sources, such as industrial effluent/laboratory-prepared solution (9), hospital effluent/laboratory-prepared solution (2), industrial effluent/other sources (1), industrial effluent/municipal wastewater treatment plant (WWTP) (1), and municipal WWTP/laboratory-prepared solution (2).
Most researchers have used Fenton for the degradation of industrial compounds (708; 56%), followed by laboratory-prepared solutions (synthetic effluent) (245; 19%), and other specific sources (98; 7.7%), such as sewage treatment plants (30; 2.4%), hospital sources (31; 2.5%), and research or educational institutions (3; 0.23%).
The volume of studies also correlates with the evolution and growth in the use of industrial processes and technologies over time, substantially increasing from the 2000s onward (Figure 1). Additionally, the search for new tools for sustainable industrial development involves introducing less environmentally harmful disposal techniques and then improving these techniques.

3.2. Size and Reaction Conditions of Fenton Reaction Studies

Most studies on Fenton reactions have been conducted on the laboratory-bench scale, representing 98.3% of the studies analyzed, followed by those on the pilot and industrial scales. On the laboratory-bench scale, a smaller volume of effluent is used under controlled laboratory conditions, even though these effluents are representative of industrial effluents, as shown in Figure 2.
Only 1.3% of the analyzed publications were conducted on the pilot scale, which is a method used for limiting the problems experienced with laboratory-bench-scale studies. However, pilot-scale studies are hindered by the operational costs associated with their implementation and execution. For example, Ferreira et al., (2020) [48] discussed the effect in a pilot plant process combining Fenton reaction with other AOPs and, then, the results showed that these kind of combinations are promising options for reducing the water footprint of biorecalcitrant industrial wastewaters. Moreover, 0.07% of direct studies have been conducted on the industrial scale, 0.07% of studies have used a combination of scales, and 0.31% of the studies did not specify the scale used.
Most studies (80.9%) used traditional reaction conditions (1022), followed by Fenton-like reactions (140) (Figure 3).
In the first case in Figure 3 on the y-axis, under ideal and optimized conditions, the compounds to be treated are catalytically oxidized in the presence of iron salts and hydrogen peroxide, in a pH range of 3.0 to 4.0. At values higher than 4.0, the rate of decomposition of hydrogen peroxide into water and oxygen markedly increases and the concentration of free iron species decreases due to the formation of iron complexes and precipitation of iron hydroxides. In the case of Fenton-like reactions, the catalyst can be other metals or complexes different from ferrous ion (Fe2+). Various monovalent and divalent cations and complexes have been considered, as noted by Jonshon and Mehrvar (2022). We found that 0.87% of articles reported the use of Fe2+ ions from secondary sources, such as industrial WWTP and mining sludge, whereas 1.3% of articles reported the use of Fe nanoparticles. The type of iron (hydro)oxide, whether natural or synthetic, solid, ionic, complex, or nano, affects the mechanism of OH formation. We identified this information in this review; according to Figure 3, the use of Fenton-like reactions with cation exchange, nanoparticles, and the combination of these conditions in the same treatment has been increasing. Pereira et al. (2012) [49] discussed catalytic forms of iron oxide [50], as well as other metal nanoparticles [51] and complex ions [52].
For pH values below 3.0, the formation of hydroxyl radicals decreases, which becomes nonexistent at pH 1.0 [53]. Under these standard conditions, the reaction time is shorter and, consequently, the removal of compounds is lower compared with those of other AOPs [54]. Notably, only under acidic conditions does the OH radical dominate, which favors the chemical oxidation process. Moreover, under normal reaction conditions, a certain amount of flocculated material accumulates, which is subsequently removed by decantation, due to the formation and complexation of iron salts [55] and the formation of highly valent oxidizing species, such as FeO3+ (Fe (V)) and the ferryl ion, FeO2+ (Fe (IV)) [56]. The concentration of hydrogen peroxide in the reaction must also be evaluated, as the excessive use of hydrogen peroxide in the reaction medium acts as a scavenger of hydroxyl radicals. The concentration of iron ions is also important for the process, ranging from 3 to 15 mg/L as, under the above conditions, the scavenging of hydroxyl radicals occurs, which decreases the efficiency of the reaction [57].
In a single study, different reactive means can be combined with the aim of comparing their efficacy. In this survey, we found that in 3.5% of the cases, the efficiencies of the traditional Fenton reaction, Fenton-like reactions, and parallel-shaped nanoparticles were tested in comparison with that of a single process.
Regarding the amount of organic matter removed, in the majority of studies, 884, more than 50% of the COD and TOC, was removed. However, only 30 studies reported that the Fenton treatment removed less than 50% of the COD and/or TOC, which was also reported in the literature [58]. For 349 studies, the reduction in COD and/or TOC was not reported.

3.3. Isolated or Combined Processes with Fenton and Electro-Fenton

Based on the articles analyzed, we collected information regarding the Fenton process being performed alone or in combination, accounting for 68.9% of the studies. Among these, 26.5% described the use of the Fenton process as the sole treatment [59,60], whereas 41.6% of the studies included Fenton combined with a second or third treatment [31,61]. Other techniques used for removing pollutants included electro-Fenton (EF) [10,62], photo-Fenton (Fe2+/UV/H2O2) [63,64], post-biological treatment, pretreatment for subsequent physical–chemical treatment, ozonolysis (UV/O3), photocatalysis (TiO2/UV/H2O2), photolysis (UV/H2O2), catalysis, UV treatment, electrocoagulation, chemical coagulation, and membranes. The removal rate was increased when the Fenton process was combined with another chemical degradation technique [65].
Using the same search methodology, we observed that 20.9% of the studies focused on the EF process, with 13.6% of the total using EF as the sole treatment process [62,66,67]. Furthermore, studies on the combination of the EF process with other AOPs accounted for 6.9% of the total [59,68]. In contrast, the results for the use of the Fenton technique showed a different scenario. Fenton has been commonly used as the primary treatment process in various studies due to its reaction design, whereas EF is more restricted in its application, as it requires additional equipment such as reactors, electrodes, and power sources [18,62].
The fastest and most thorough destruction of target organic pollutants in water is currently achieved using Fenton-based electrochemical advanced oxidation methods [9]. The original method was EF; however, thanks to the occurrence of important photoreduction approaches, UV and solar photo-EF processes achieve a greater mineralization [9,53,62]. In actuality, the two-electron oxygen reduction reaction taking place at the cathode becomes ineffective or insufficient as solution pH rises, limiting the decontamination’s efficacy [9,53,62]. Because of this, recent research has been concentrated on two key aspects of the EF processes: (i) developments in cathodic H2O2 electrogeneration, which demonstrate how the oxygen reduction reaction improves upon use of new and/or more sustainable electrocatalysts, cathode configurations, and reactor designs; and (ii) improvements in iron-based catalysts, with the primary goal of extending the application to a much wider pH range and ultimately surpassing the classical acidic limitation [9,53,62,69].
The electrochemical production of H2O2 is the basis of the Fenton-based EAOPs. Although the cathodic approach using the two-electron oxygen reduction process (Equation (1)) is considerably more common, this can also be created anodically [9,10,53]
O2 + 2H+ + 2e → H2O2
After then, H2O2 can be either dropped in the medium to be progressively degraded, transferred to the anode to be oxidized to O2 via the hydroperoxyl radical (HO2), or reduced cathodically to water [9,53,69]. The obtained cumulative H2O2 concentration depends on the cell configuration, the kind of cathode, and a number of operational factors. After that, H2O2 can be reduced cathodically to water, transported to the anode to be oxidized to O2 by the hydroperoxyl radical (HO2), or left in the medium to gradually degrade. The cell configuration, the type of cathode, and several operational parameters all affect the cumulative H2O2 concentration that is achieved [9,53].
Despite the fact that H2O2 is a weak oxidant, the Fe2+ ion promotes the occurrence of Fenton’s reaction (Equation (2)). The greatest formation of free OH in the bulk is achieved at its ideal pH, which is close to 3. Because the resultant Fe3+ may be continually reduced to Fe2+ at the cathode surface via reaction (Equation (3)), continuing the catalytic Fe3+/Fe2+ cycle, this so-called EF process is more effective than the traditional Fenton [9,10,53,69].
Fe2+ + H2O2 → Fe3+ + OH + OH
Fe3+ + e → Fe2+
PEF uses a UVA lamp to irradiate the solution as a photo-assisted method of EF. Reaction theory suggests that this increases the amount of OH produced as a result of the photolysis of [Fe(OH)]2+, the main Fe3+ species at pH ~ 3 (Equation (4)) [9,53]. Moreover, the UVA photons accelerate the mineralization of the contaminated solution by causing the rapid photodecomposition of the final Fe(III)-carboxylate complexes via the general reaction (Equation (5)).
[Fe(OH)]2+ + hν → Fe2+ + OH
[Fe(OOCR)]2+ + hν → Fe2+ + CO2 + R
Instead, a UVC lamp can be employed, with the goal of producing more OH through the photolytic homolysis of H2O2. Moreover, the aromatic contaminants may be photolyzed by the UVC irradiation, hastening their degradation [9,53]. Using the SPEF method, which uses free sunlight as the energy source, can fully eliminate the energy consumption associated with the use of expensive UV lamps, increasing the viability of the effective photo-assisted EF process. The substantially increased light power given by sunshine clearly outperforms the PEF method [9,53,69]. The experimental setups for these processes consist, essentially, of undivided two-electrode reactors with an air-diffusion cathode. For this reason, different oxidants produced at the anode can concurrently oxidize organic contaminants in an undivided cell [9,53]. While active chlorine (Cl2/HClO/ClO) may also be formed by Cl oxidation in chloride medium, adsorbed M(OH) is the most prevalent one in aqueous media. The primary oxidizing agents in anodic oxidation are M(OH) and/or active chlorine, which may improve EF’s capacity to degrade organics in the water matrix. The most relevant anodes for producing oxidants via the anodic approach are thought to be synthetic diamond electrodes [9,53]. Nevertheless, it is important to consider that the nature of the anodic material has a less significant role in the PEF and SPEF treatments, where the photons’ action on the removal of the pollutants is more vital. Thus, PEF or SPEF pilot plants may employ substantially less expensive materials, such as dimensionally stable anodes [9,10,53,69]. Frequently, the total organic carbon decay and/or chemical oxygen demand of treated solutions indicate a significant efficacy of mineralization [9,53]. In this frame, the mineralization current efficiency, apparent current efficiency, and energy consumption demonstrate the treatment’s feasibility. But, an increase in current accelerates the degradation and mineralization of the contaminants due to the quicker electrochemical synthesis of oxidants, although this effect is attended with a decrease in current efficiency and an increase in energy consumption due to the parasitic reactions [9,53,69].
It is possible to employ divided cells to create H2O2 in the catholyte, which is then kept for injection into a reactor where chemical Fenton treatment is then applied. Undivided cells are still the best choice to reduce cell voltage and save money on membranes [9,53,62]. For H2O2 electrogeneration in Fenton-based EAOPs via reaction (Equation (1)), there are traditionally two major cathode topologies to be distinguished [9,53]: (i) large surface area immersed/soaked porous structures or particle beds where the dissolved O2 (from sparged O2 or air) is reduced; and (ii) gas-diffusion electrodes (GDEs) formed by a gas-diffusion layer through which pumped atmospheric air or pure O2 flows from an air chamber and a catalyst layer (i.e., a mixture of polymer with raw or modified carbon) where reaction (1) takes place [9,53,70]. When the gas encounters the solid and liquid phases at the three-phase boundary (TPB) [9,53,71], H2O2 is produced in both situations. The solubility of this gas restricts the yield when the former type is utilized, whereas the GDEs tend to be more effective and result in higher H2O2 concentrations due to the infinite supply of O2 [9,72]. Other authors have merged the two designs to create a GDE that resembles three dimensions and is made of a hydrophobic carbon cloth on top of a porous graphite felt, achieving mineralization current efficiencies greater than 100% in SPEF [9,73]. Recently, a new version has been created that is theoretically more like the GDE [9,53,62]. It is made up of a floating (also known as air-breathing) cathode that is partially exposed to the environment to allow for natural air diffusion, negating the need for an air pump [53,74]. These cathodes are superhydrophobic (superaerophilic, in other words), and by increasing O2 diffusion, they enhance the oxygen reductive reaction (ORR) catalytic performance [10,75]. By varying the polymer composition, air-breathing cathodes’ hydrophilicity and hydrophobicity may also be adjusted, as can the catalyst’s porosity. Zhang et al. (2020) [76] investigated the impact of cathode location on the generation of H2O2 revealing a higher current efficiency than a GDE. Recent work has focused on scaling up floating cathodes, which resulted in an H2O2 generation rate of 196.3 mg h−1 [9,53,77]. It is important to note that there are alternative methods for also improving the feasibility of cathodic H2O2 electrogeneration for usage in EAOPs [9,10]. They are based on innovative reactor designs created for self-sustaining H2O2 generation, eliminating the need for the air pump [9,53]. Pérez et al. (2016) [78] have shown that a jet aerator based on the venturi (i.e., suction) action induced by a solution running at high velocity close to a tiny air intake exposed to the environment can produce enough H2O2. Scialdone et al. (2015) [79] and Hamdi et al. (2020) [80] suggested using pressured EF systems to supersaturate the solution with O2 to decompose 8-hydroxyquinoline-5-sulfonic acid and Acid Orange 7 dye. One order of magnitude more H2O2 was produced during ORR at a graphite cathode using air compressed at 11 bar than at ambient pressure [9,79]. It has also been tried to combine the two reactor types mentioned, creating a pressurized jet-aerator. In another study, the same authors have also suggested a different reactor design to considerably boost the generation of H2O2 during the EF process. The H2O2 electrogeneration was improved by an order of magnitude using a microfluidic electrochemical device, which had an ideal nominal distance of 120 µm between the cathode (an inexpensive graphite plate) and the anode [81]. An alternative strategy is increasing the mass transfer of naturally dissolved oxygen to the cathode surface, for instance, using high turbulence promoters such as a revolving cylinder electrode as the cathode [82]. Another bright idea, tested with various anode–cathode configurations and liquid flow rates, attempts to use the anodically generated O2 for further H2O2 synthesis in flow-through reactors [83,84].
The electrocatalysts typically demonstrate a trade-off between activity and selectivity, making it simple to make a decision that balances the amount of time needed to treat the water with the efficiency of the process (closely related to the H2O2 generation efficiency) [9,53,62]. It may be possible to simultaneously boost activity and selectivity for the formation of H2O2 by using heterogeneous single-atom catalysts with atomically scattered metal atoms on 2D materials [9,53,85]. An excellent illustration of this is a single Zn-centered phthalocyanine; however, the N ligands are gradually destroyed by H2O2, as is the case with porphyrins containing 3d transition metals.
Meanwhile, the kinetic current density of a Co-N4 moiety inserted in N-doped graphene was increased by fine-tuning the atomic structure, with no activity loss for 110 h [86]. Yet, a significant shortcoming of single-atom catalysts continues to be their instability. Noble-metal-based electrocatalysts are also an excellent option for producing H2O2 because of their high activity, selectivity, and stability. There are also other options, such as Au-Pd [87] and Hg alloys with Pt, Ag, and Pd [88]. However, their complex manufacture, high expense, and usage of hazardous starting materials create a bottleneck for their actual use in water treatment. They may be ideal for producing H2O2 in small devices [9,53].
Carbonaceous cathodes are ideal for use in EF and related EAOPs [9,53]. Carbon catalysts are readily available, affordable, nontoxic, durable, and may be extremely reaction-selective (Equation (1)) [9,89] and the beneficial effects of surface imperfections, particularly the edges, have been experimentally demonstrated [9,10,53,90]. Although novel varieties of carbon-based and carbon-supported electrocatalysts have recently emerged, immersed carbon felt [9,53,91] and GDEs manufactured with carbon black (CB) particles combined with PTFE [9,10,53,92] are still the most widely used traditional carbonaceous cathodes for EF processes. As a strategy to boost activity, several scientists have adopted the switch from nanoparticles (like CB) to high-surface-area nanostructures. For instance, mesoporous carbons, CNTs, and graphene have all been employed in the EF process [9,53,93,94,95]. Macroporous (i.e., three-dimensional) carbonaceous cathodes are also widely used in order to support reaction (3), which would lead to continuous Fe2+ electroregeneration [9,10,53]. Intentionally created fissures have also been used to create porosity [96]. Undoped and F-doped hierarchically porous carbon (HPC) have been recently created by Zhao et al. (2018) [97] as new structures with strong electrocatalytic activity to generate H2O2. These materials are created by the carbonization of metal–organic frameworks (MOFs), which produce an abundance of microporous, mesoporous, and macroporous materials rich in catalytic sites and improve the kinetics of ORR [98].
Several researchers have tried to modify carbons using various methods to improve their selectivity using organic solvents as well as chemical and electrochemical methods [99] and air calcination [100]. These modifications increase the cathode’s hydrophilicity and electroactivity by causing oxygen and nitrogen functionalities [9,10,53]. For example, the electro-oxidation of graphite felt led to a sharp rise in H2O2 concentration from 3.2 mg L−1 with the raw felt to 37.5 mg L−1 at 90 min [99]. Meanwhile, the S doping of mesoporous carbon achieved a selectivity for H2O2 of about 80% [101], the P doping of carbon nanotubes [102], or the addition of B in addition to the traditional O and N [103]. Also, surface modification using quinones can be achieved, for example, with azo dyes, or Co-phthalocyanine [104]. The greatest concentration of H2O2, achieved at 100 mA cm−2 (after 90 min of electrolysis in 0.1 M K2SO4 at pH 3.0), increased from 770.7 to 1025.9 mg L−1 due to the inclusion of the azo dye [10,105]. Alternately, carbon can also become more selective when it is decorated with a variety of particle kinds. A number of chalcogenides have so far been explored, and co-structures are particularly well-suited for the generation of H2O2. Due to the great selectivity (80%) of CoS2 with pyrite structure, GDEs built from decorated carbon nanotubes on carbon cloth or paper are now possible to produce [106]. Bronopol and bentazon can be degraded by EF and PEF since it has been demonstrated that the presence of phosphorus atoms in substitutional locations imparts a significantly stronger stability [107,108] as well as the particles made of W@Au [109] or NiFe produced from MOFs [110] have been employed. The recently discovered superhydrophobic cathodes are one example of how wettability/hydrophobicity is important. It can be possible by mixing silicon or a polymer with carbon to help trap more O2 molecules, particularly in submerged cathodes. This effect has been demonstrated with the use of a breathable dimethyl silicon oil maturation layer, which lowers water flooding and increases rates of O2 transfer, thus improving electrogeneration activity [10,111]. In contrast to cathodes that are exclusively hydrophobic or hydrophilic, a Janus cathode made up of a hydrophobic gas-diffusion layer and a hydrophilic catalyst layer was recently reported to increase the generation of H2O2, reducing the cost of the process [112].
It is important to note that one of the major trends in recent years has been the production of sustainable carbons, which has led to the development of greener electrocatalysts made from biomass [113]. According to the existing literature, these materials have mostly been examined for their ability to produce H2O2, utilizing bean deposits [114], glucose [115], or fungal hyphae [116]. Contrarily, Fenton-based EAOPs occasionally need cathodes, as was recently demonstrated for carbons generated from chitosan in the EF treatment of acebutolol [117].
Because acid pH is ideal for Fenton’s reaction, the cathodes and reactors described in the preceding section have often been employed to treat polluted effluents via EF, PEF, and SPEF [118]. Yet, many industrial effluents and natural water have pH levels that are close to being neutral, which makes it difficult to apply traditional homogenous techniques. Two approaches may be used to overcome this disadvantage [119]: (i) heterogeneous catalysis using various types of solid iron, and (ii) homogeneous catalysis using soluble Fe(III) complexes (i.e., chelated Fe(III)). To create complexes with Fe, short-chain aliphatic carboxylic acids, like oxalic acid and citric acid, have been employed in the past. Due to their higher photoactivity, which enables a higher Fe2+ regeneration, aminopolycarboxylic acids like ethylenediaminetetraacetic, nitrilotriacetic, and ethylenediamine-N,N′-disuccinic (EDDS) have recently been proposed as possible chelators. Due to its greater biodegradability, EDDS appears to be a preferable option among them. The continual electroreduction of the complex at a carbon–felt cathode surface, which is preferable than a GDE, has been credited with the high performance of the EF-like process using the Fe(III)-EDDS complex as catalyst [120]. The photoirradiation in a PEF system is an option to assure the production of Fe(II) [121]. All of these instances involve the progressive destruction of the organic chelator by OH, even if the target pollutants can already be completely degraded before this happens [62]. It is possible to conduct heterogeneous Fenton-based EAOPs using natural minerals such as chalcopyrite and pyrite (FeS2) [10,91,122,123]. For Fenton or EF treatments, crystalline pyrite is a particularly strong option since it is a superb electron donor whose S22− conversion to sulfate is followed by the release of Fe2+ and H+. Due to this, two degradation pathways—conventional Fenton’s reaction, whose occurrence is encouraged by progressive acidification, and heterogeneous Fenton’s reaction—can coexist. Nevertheless, considerable iron leaching restricts the pyrite-catalyzed EF’s capacity to be reused and necessitates sludge management, despite the fact that it has demonstrated higher efficacy than other heterogeneous EF treatments [62]. In the case of vermiculite for dye treatment, iron-rich soils such as clays have also been researched [124]. Another source of solid iron is found in artificial iron-loaded materials, such as fly ash, modified cathodes, resins, zeolites, membranes, biosorbents, and hydrogels. Additional artificial structures include Fe-MOFs and Fe-MOF-derived carbons, which are potentially extremely porous nano/microcatalysts, and nanozerovalent ions [10,125,126,127,128,129].
A porous carbon modified with nanosized zerovalent iron (nano-ZVI) produced during pyrolysis of a 3D MOF was used. Its magnetic characteristics enabled recovery at the conclusion of the treatment, and it also demonstrated a high level of stability as shown by the little iron leaching and good recyclability. The utilization of an appropriately little amount (0.05 g L−1) of a non-pyrolyzed 2D MOF in PEF treatment at pH 7.4 was granted. In this instance, adding carboxylic groups to the ligand improved the raw 2D MOF’s resistance to hydrolysis [10,129].

3.4. Journals Publishing on Fenton Indexed in SCOPUS through 2022

The articles about Fenton were published in a total of 400 journals. The distribution of these publications was asymmetrical; Table 2 displays the 23 journals that had the highest number of publications in descending order, representing a total of 49.2% of the total publications.
To assess the quality of the articles and their publication venues, we conducted an evaluation using the Journal Citation Indicator (JCI). The JCI is a metric that is widely used to calculate the impact of a citation between articles and reviews published in a journal, considering the Journal Impact Factor, and reflecting the journal’s performance over the past 3 years.
In this study, among the top 10 journals, four had a JCI above 1 (considered above average for their category): Journal of Hazardous Materials (1.94), Chemosphere (1.48), Journal of Environmental Management (1.38), and Separation and Purification Technology (1.43). Chemical Engineering Journal (2.13) stands out, with a JCI above 2, indicating that its performance was more than twice the average in its category.
In terms of the total number of publications (1263), the journals Desalination and Water Treatment (68), Chinese Journal of Environmental Engineering (67), and Journal of Hazardous Materials (66) were the most prominent, publishing a total of 15.91% (201) of the articles considered in the topic of this bibliometric analysis.
When considering the country of the corresponding author, the studies included in this review of Fenton research encompassed authors from 66 different countries. The distribution of publications among countries is uneven, with 10 countries accounting for approximately 75% of the total number of publications (as shown in Table 3). China produced the largest number of Fenton-related studies (530; 41.9%), followed by Turkey (95; 7.5%) and Iran (86; 6.8%). Meanwhile, Brazil ranked eighth with 29 published articles, representing 2.29% of the total.
In terms of research and sustainable economic policies, China has established itself as a leader in the field, as reflected in its Scientific Electronic Library Online (2013) ranking. China is at the forefront of scientific-article production in the developing world, having experienced substantial growth in this area mainly owing to its exponential technological and industrial progress [130].
Additionally, several European countries, including Spain, Portugal, Poland, and Italy, are leading in industrial waste treatment within the bloc. This is due to the European Union’s increased enforcement of environmental legislation, which has resulted in a higher number of scientific publications. Spain is a notable country on this list, with legislation based on the European Framework that covers the treatment of liquid and solid waste at all levels, strictly regulating waste and setting high standards [131]. Turkey has also achieved progress in industrial waste treatment owing to government support for projects focused on treating industrial effluents and recycling solid waste. Several Asian countries feature on the list of the top 10 countries publishing scientific research on waste treatment. Iran, India, and Taiwan can be noted for their strong regulatory environmental legislation [132]. In Brazil, which ranks eighth on the list, waste treatment has become a priority in recent years, and a consolidated legal framework is increasingly necessary to ensure compliance with environmental regulations and permits [133].

3.5. Perspectives on Water Reclamation after Fenton Treatment

The importance of water as a vital resource for humanity cannot be stressed at this difficult time of climate change. The annual demand for water is expected to increase because of the growing challenges that the world’s population faces. According to estimates, by 2030, the global water demand would be 6900 billion m3, or around 64% more than what most countries now have access to [134]. It is essential that the world prioritizes increasing the availability of clean water through alternative techniques, such as desalination, and water recycling and reuse. In the case of the treatment processes based on the Fenton reaction, these techniques have been effectively applied to treat various types of industrial effluents and municipal sewage-treatment plants in order to reduce the load of refractory pollutants. However, particular attention should be given to the side effects of the treatment, such as excesses of soluble ferrous ions.
Therefore, to ensure the availability and sustainable management of water and sanitation for all, water reuse and recycling as well as water reclamation after Fenton-based processes are essential strategies to accomplish SDG 6.
According to Domingues et al. (2022) [135], the total iron content accepted in Portuguese legislation is 2 mg L−1. Therefore, several alternatives for the reduction of iron ions in the water matrices are being studied. For example, Volesky (2001) [136] has investigated the use of biomass, or Martins and co-workers (2017) [137] have studied the use of ion exchange resin. Both strategies can be steps in the Fenton post-treatment to suit the safety standards of potable water. It is worth mentioning that the possibility of Fenton-like reactions has been one of the most discussed reactions today [18], associated or not with the steps of reduction and adaptation to the proposed use of treated water [138]. In comparison to traditional Fenton procedures, the fluidized-bed Fenton (FBF) process offers a superior way for decreasing the creation of significant amounts of iron oxide sludge. It is frequently followed by the coagulation sedimentation (CS) process. In order to remove micro-pollutants and their by-products, FBF has found extensive use in the treatment of industrial wastewaters, including those from pharmaceutical, pesticide, textile, dye, and other refractory organic sources [139,140,141]. As a further low-cost and energy-efficient unit procedure, constructed wetlands (CW) have been suggested. According to Arden and Ma (2018) [142] and Liu et al. (2019) [143], CW systems integrate a wide variety of biotic and abiotic processes to offer a green solution for a variety of treatment applications. Before wastewater from wastewater treatment plants (WWTPs) is discharged into rivers, CW are frequently used to further treat the pollution [144]. This comprises effluents from many sources, such as home wastewater-treatment facilities [145], freshwater aquaculture facilities [146], the textile sector [147], and fertilizer-production facilities [148]. Previous investigations have demonstrated the effectiveness of CW in removing several pollutants and reducing important parameters, including chemical oxygen demand (COD, 78–81%), biological oxygen demand (BOD, 72–82%), total solid suspension (TSS, 79–89%), nitrogen (84%), color (74%), phosphorus (79%) [145,146,147], total organic carbon (TOC, 98.5%) [149], chlorpyrifos (94–98%) [150], organophosphorus pesticides (87.22 ± 16.61%) [143], steroid hormones (97.4 ± 0.09%), and biocides (92.4 ± 0.54%) [54]. Therefore, it is important to couple this strategy with Fenton-based technologies for real applications.
A typical urban river with little natural discharge (less than 5%) is a real example of this coupled strategy. The Shuangji River is found in the northern Chinese province of Henan. The greatest recycled paper manufacturing facilities in China are in the Shuangji River basin. Papermaking WWTPs dump approximately 75% of the river’s recharge water, which severely pollutes the river’s water system and causes the demise of aquatic species. A WWTP with CW was built with a daily capacity of 120,000 tons of wastewater treatment to address this problem. Then, FBF, CS and then the application of CW are all used in this integrated process, which has proven to be successful in lowering toxicity and raising water quality, eliminating the Fenton residues and, consequently, the water scarcity can be effectively addressed by using wastewater from the pulp and paper industries as reclaimed water. These conclusions were based on laboratory-scale and large-scale experiments, achieving under-optimal conditions for FBF treatment (pH (3.5), H2O2 dosage (0.93 mL/L), H2O2/Fe2+ ratio (4), and hydraulic retention time (HRT, 60 min)), achieving average removal efficiencies of about 87.3%, 93.59%, 51.73%, 84.75%, and 95.86%, for COD, ammonia nitrogen (NH3-N), total nitrogen (TN), total phosphorus (TP), and color, respectively. The amount of the genotoxic substance 4-nitroquinoline 1-oxide (4-NQO-EQ) was reduced from 30.6 ± 1.6 g/L in the influent to 12.4 ± 1.0 g/L after treatment by FBF, and then it was further reduced to 5.9 ± 0.4 g/L after treatment by CW and finally reached 3.2 ± 0.3 g/L after self-purification downstream for 12 km. Zebrafish chronic survival rates considerably increased over a 21-day period, rising from 0.0% in the influent to 58.8 ± 4.0% in the CW effluent before progressively increasing to 68.8 ± 2.6% after the 12 km downstream self-purification. Through this, the concentration of 8-hydroxy-2-deoxyguanosine in zebrafish dropped from 120.0 ± 19.3 ng/L in the ecological oxidation pond effluent to 94.0 ± 7.5 ng/L in the effluent of CW, and even further to 42.0 ± 3.0 ng/L after the 12 km downstream self-purification. In conclusion, the FBF-CW treatment method successfully detoxified wastewater from papermaking while also improving the quality of the water, making it appropriate for use as an ecological water supplement [151].
FBF techniques have gained popularity in recent years for the treatment of difficult industrial effluent. The advantages of sludge reduction provided by heterogeneous Fenton processes are combined with the efficiency of homogeneous Fenton reactions in FBF. FBF exhibits more promise in terms of cost-effectiveness and scalability when compared to other modified Fenton processes. However, the high Fenton reagent consumption rate and the requirement for exact pH control continue to be barriers to the full-scale application of FBF. Therefore, a classification for FBF processes has been developed into homogeneous, and heterogeneous approaches based on their response mechanisms and system designs as well as modeling methods like computational fluid dynamics models and artificial neural networks have been investigated because these have the potential to accelerate the full-scale adoption of FBF technology [152].
Another way to intensify the removal of organic matter is the combination of Fenton treatment with chemical coagulation techniques. For that, Metin and Cifci (2023) [153] reported that COD removal removed about 92%, and 96% of total suspended solids when a real effluent was treated applying this novel treatment strategy. In terms of cost, these authors determined that this combined treatment has a lower cost ($19.16 m−3) when compared to treatment with photo-Fenton using UV irradiation.
The precipitated iron sludge can be affected by the environment of the reaction solution causing a negative impact on the practical application of Fenton technology. Iron catalyst deactivation, dosage amount additions and sludge-management issues [118] are also disadvantages on the use of the Fenton applicability. However, Fe3+ is also a well-known flocculant, and some investigations have indicated that the hydrolysis of Fe3+ also contributes to the removal of pollutants in the Fenton process. During the hydrolysis process, some pollutants can be removed by the coagulation and adsorption of Fe3+ [154]. The homogeneous Fenton oxidation process followed by a pH neutralization to accelerate coagulation for the treatment of landfill leachate was utilized and 61% of COD was removed and the contribution ratio of Fenton oxidation and coagulation was 0.75 [154]. Meanwhile, a pesticide wastewater using the oxidation–coagulation process in the Fenton process was also treated, reducing the COD from 33,700 to 9300 mg L−1, and, consequently, resulting in a significant improvement in biodegradability [154]. In addition to coagulation, iron salt can also play a role of adsorption to remove pollutants when these are hydrolyzed. The in situ-formed nano-hydrolyzed Fe3+ particles (nano-FeOOH) have a higher specific surface area and can adsorb the carboxyl-rich intermediate product generated by Fenton oxidation [155]. The above results show that Fenton is a synergistic process of oxidation, coagulation, and adsorption. Normally, it is difficult to achieve the discharge standards of complex and refractory wastewater through a single treatment process. Then, multiple combined processes including adsorption, coagulation and advanced oxidation are often required.
Even though they are still made, direct olive mill effluents (OMW) are discharged into surface waters, which is dangerous, against the law, and has serious contamination effects. Thus, it is not acceptable to use OMW for irrigation [156,157]. According to several studies [158,159,160,161], the discharge of OMW has negative effects on the aquatic fauna and ecological balance as well as odor problems, soil contamination, plant-growth inhibition, underground leaks, water pollution, and disruption of self-purification processes [162]. For example, due to the significant amounts of pollution found in the Guadalquivir River Basin, the Spanish government outlawed the direct release of OMW into rivers in 1981. As a general remedy for OMW’s natural evaporation, artificial lagoons were built in nations like Spain. The low evaporation rates, odor emissions, and dangerous leaks brought on by poor construction have, however, shown over time that this strategy is ineffective. Thus, chemical remediation techniques are required to clean up these resistant wastewaters [155,156,157,163,164,165]. In an olive oil mill in Jaén, Spain, which employs a cutting-edge two-phase olive-oil-extraction technology, earlier studies have improved a treatment process based on Fenton’s reaction for the reclamation of OMW. The primary treatment (PT), which removes coarse particles, is followed by advanced oxidation with Fenton-like characteristics, flocculation–sedimentation, and filtration through olive stones in series (ST) [164,166,167,168,169]. However, the measured electroconductivity (EC) values in the effluent at the ST outlet (OMW/ST) are higher than the advised range of 2–3 mS cm−1, indicating elevated salinity levels that pose a risk in accordance with the guidelines established by the Food and Agricultural Association (FAO) for using regenerated water for irrigation [170].
In recent years, membrane technology has also become a very attractive alternative to conventional separation methods, offering a number of advantages [171,172,173]. Although membrane technology has advanced significantly over the past 50 years, much work still needs to be done to improve its competitiveness in large-scale industrial applications. Membrane fouling is one of the key issues with membrane technologies, especially in wastewater-purification operations [162,174,175,176,177,178]. Fouling reduces flux quickly and jeopardizes the membrane’s overall effectiveness. Due to the increased operational and energy expenses caused by this decrease in productivity, plants must frequently shut down for in situ membrane cleaning. Unfortunately, these cleaning techniques fall short of fully restoring the membrane module’s initial effectiveness, leading to a gradual decline over time. Prior research on the reuse of OMW utilizing membrane technology has frequently noted problems such as concentration polarization and fouling formation on the membranes. These issues lead to a substantial loss of flux and permeate quality, which raises energy costs and makes the operation unprofitable. The selection of the wrong membrane types and insufficient operation conditions are the main causes of these fouling problems. In the study carried out by Ochando-Pulido and co-workers (2012) [162], the physicochemical properties of OMW after ST, which entails flocculation, filtering in series, and a Fenton-like reaction were investigated to determine the best membrane and avoid fouling during following pressure-driven membrane operations for the effluent’s complete purification [162]. Analyzing bacterial growth, gauging saturation index, figuring out particle size distribution, and gauging the distribution of organic matter molecular weight cut-off (MWCO), the OMW ST was effective in reducing the dangerous electroconductivity (EC) values (2–3 mS cm−1) and determining that the significant amount of organic contaminants (31.7%) with average diameters less than 3 kDa are relevant for membrane fouling. The saturation index suggested working with a recovery factor of less than 90%. Finally, using the chosen NF membrane at a pressure of 15 bar guaranteed low fouling, high flux generation (69.9 L h−1 m−2), and significant rejection efficiency (55.5% and 88.5% for EC and COD). This makes it possible to obtain an effluent of high quality in accordance with the FAO requirements in order to reuse the recycled water for irrigation.
Another study proposed a unique technology called advanced oxidation–nanofiltration hybrid technology in keeping with the developing circular economy concept. The method was created through experimental research on wastewater samples collected from a tannery factory in Kolkata [179]. A unique membrane made of graphene oxide (GO) is created especially for the nanofiltration stage, and the effectiveness of this membrane is contrasted with those of other commercially available membranes used to eliminate the residues after Fenton-based process. The reuse of the purified water was possible by subsequent microfiltration and subsequent nanofiltration. The synthetic GO-based membrane functioned remarkably well, achieving significant COD (99%), TDS (>96%), and chromium (>99%) removals. These findings demonstrate the promise of the novel GO-based membrane-equipped advanced oxidation–nanofiltration system as an efficient and sustainable method for treating wastewater from the leather industry, enabling its recycling and reuse [179].
To lessen the concentration of persistent pollutants, several industrial effluents and municipal sewage have been successfully treated using Fenton reaction-based processes. The presence of higher amounts of soluble ferrous ions, an issue in the wastewater treatment sector, is another one of the treatments’ potential adverse effects that must be taken into account. According to studies by Domingues et al. (2022) [135], the highest allowable limit for the total iron concentration in wastewater released into natural water bodies or used as a source of public water in Portugal is 2.0 mg/L [135].
In a study by Victor-Ortega and co-workers [180], it was determined if Dowex Marathon C resin, a strong-acid cation resin, was effective at removing leftover iron from the supernatant liquid after iron precipitated during Fenton’s oxidation of olive mill effluent (OMW). Their research concentrated on iron levels that were rather modest, with feed concentrations reaching up to 5 mg/L. According to the authors, the final effluent had an iron concentration of 200 g/L, which is less than the allowable limit for reusing water. In other related studies by the same authors, sodium, chloride, iron, and phenols were removed from a simulated OMW that underwent a chemical treatment involving Fenton’s peroxidation and iron precipitation using Dowex Marathon C (strong-acid cation) resin and Amberlite IRA-67 (weak-base anion) resin.
A study on the efficacy of the Fenton reaction in treating wastewater from the paper industry found that the removal efficiencies for TSS, BOD, COD, TKN, and total phosphorus (TP) were all very high, with values of 95.4%, 92.7%, 93.7%, 90.5%, and 91.7%, respectively. This study showed Fenton treatment’s effectiveness in permitting the reuse of paper industry wastewater for irrigation while meeting Egyptian code irrigation criteria [181]. The goal of the current study was to increase greywater treatment’s effectiveness by using a chemical-treatment strategy. In order to decrease the physicochemical properties of the greywater to levels appropriate for irrigation, as determined by the applicable regulations, this chemical treatment involved the employment of Fenton’s reaction as an advanced oxidation process.
According to Ribeiro et al. (2016) [182], clothes colored using wastewater treated with Fenton’s reagent (reused water) produce results that are comparable to those of fabrics dyed using conventional dyeing techniques (production waters). When it came to the pink and gray fabric samples, Fenton’s wastewater treatment produced even better results than reclaimed water, which, because of its natural hue, significantly darkened the tone of the materials. All fabric samples passed stringent industrial quality control, and the results were all completely satisfactory. The fabrics made utilizing wastewater treated with Fenton’s reagent did not exhibit any instances of stains, discoloration, opacity, or any other unfavorable impacts. These results show that wastewater that has also undergone post-treatment is suitable for reuse in the textile-dying procedure.
The discussion above makes it very evident that waste is produced by all water purification and recycling methods, especially the Fenton process. As a result, waste management is crucial for both the environment and human health. This is due to the fact that waste materials are typically poisonous. The produced waste could contaminate water supplies once more. Screening, filtering, centrifugal separation, sedimentation, gravity separation, coagulation, flotation, aerobic and anaerobic processes, evaporation, and precipitation all produce a lot of solid waste. Water with a high concentration of contaminants is produced via reverse osmosis, micro-, and ultra-filtration, and Fenton-based processes. For the management of waste materials produced during water-treatment procedures, numerous strategies have been devised and used. Recycling sludge into useful items like fertilizers, fillers, building materials, etc., is the most crucial waste-management strategy. A few hazardous waste materials have been burned, and the ash from those fires has been used as fertilizer. It has also been suggested to bury the waste materials underground in containers constructed of hermetic plastic or iron. These days, beneficial and affordable adsorbents are utilized to remove contaminants from water, including fly ash, red mud, and sand, and these can be used after Fenton treatment. Large amounts of these exhausted adsorbents have been manufactured, and they have been successfully used as fillers and building materials.

4. Conclusions

The environmental impacts caused by the waste generated by industries worldwide require attention and specific treatment that is of low cost and easy to operate, and that combines the efficient removal of recalcitrant carbon and other micropollutants, both on medium and large scales. Through this survey, we identified a substantial increase in publications focusing on the Fenton method from the 2000s onward. However, most of these publications are still concentrated in China and Europe, highlighting the need for more incentives for the treatment of alternative waste in other countries.
As suggested in the literature, the Fenton process is an effective and cost-efficient method for treating a diverse range of compounds found in various types of effluent matrices. This process particularly efficient at removing COD, with satisfactory results of over 50% achieved in most studies. The results of this review showed that one of the most common applications of the Fenton process is in treating dyes and industrial effluents in general. Another feature that must be considered is that EF is a Fenton-based technology which has received great attention in the last decade, allowing significant technological advances on the reactors and catalysts [10].
This systematic survey also demonstrated the need to conduct studies on larger operational scales, so that the results are effective and applicable in industrial settings; most studies have been conducted on the laboratory scale, lacking adequate transposition to larger scales. Thus, this review can serve as a benchmark for the initiation of new studies involving the Fenton process, helping to establish it as an important chemical process for the treatment of effluents.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/pr11082466/s1, Figure S1: Time trend for the number of articles published between 1995 and 2020 in the Web of Science database, considering compounds degraded by Fenton; Figure S2: Time trend for the number of articles published between 1995 and 2020 in the Web of Science database, considering the origin of effluents degraded by Fenton; Figure S3: Types of scale in which the Fenton studies were carried out; Figure S4: Types of matrices in which the Fenton studies were carried out.

Author Contributions

Conceptualization, M.P.P., S.M.d.O.L.M., K.B.M., D.M.d.A., C.A.M.-H., A.L.L.A. and J.E.F.d.S.G.; methodology, M.P.P., S.M.d.O.L.M., K.B.M., D.M.d.A., C.A.M.-H., A.L.L.A. and J.E.F.d.S.G.; software, M.P.P., S.M.d.O.L.M., K.B.M., D.M.d.A., A.L.L.A. and J.E.F.d.S.G.; formal analysis, M.P.P., S.M.d.O.L.M., K.B.M., D.M.d.A., A.L.L.A. and J.E.F.d.S.G.; investigation, M.P.P., S.M.d.O.L.M., K.B.M., A.L.L.A. and J.E.F.d.S.G.; resources, S.M.d.O.L.M., K.B.M., A.L.L.A. and J.E.F.d.S.G.; data curation, M.P.P., S.M.d.O.L.M., K.B.M., A.L.L.A. and J.E.F.d.S.G.; writing—original draft preparation, S.M.d.O.L.M., D.M.d.A. and C.A.M.-H.; writing—review and editing, S.M.d.O.L.M., K.B.M., D.M.d.A., C.A.M.-H., A.L.L.A. and J.E.F.d.S.G.; supervision, M.P.P., S.M.d.O.L.M., K.B.M., A.L.L.A. and J.E.F.d.S.G.; funding acquisition, S.M.d.O.L.M. All authors have read and agreed to the published version of the manuscript.

Funding

Financial support from Conselho Nacional de Desenvolvimento Científico e Tecnológico (CNPq, Brazil) [306323/2018-4, 312595/2019-0, 439344/2018-2, 315879/2021-1, 409196/2022-3, 408110/2022-8], and from Fundação de Amparo à Pesquisa do Estado de São Paulo (Brazil), [FAPESP 2014/50945-4 and 2019/13113-4] are gratefully acknowledged.

Data Availability Statement

Data will be made available on request.

Conflicts of Interest

The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.

References

  1. Boretti, A.; Rosa, L. Reassessing the projections of the World Water Development Report. Clear. Water 2019, 2, 15. [Google Scholar] [CrossRef]
  2. Khaya, P.S.; Babatunde, F.B.; Joseph, K.B. The Treatment Effect of Chemical Coagulation Process in South African Brewery Wastewater: Comparison of Polyamine and Aluminum-Chlorohydrate coagulants. Comparison of Polyamine and Aluminum-Chlorohydrate coagulants. Water 2022, 14, 2495. [Google Scholar] [CrossRef]
  3. Konradt, N.; Kuhlen, J.G.; Rohns, H.-P.; Schmitt, B.; Fischer, U.; Binder, T.; Schumacher, V.; Wagner, C.; Kamphausen, S.; Müller, U.; et al. Removal of Trace Organic Contaminants by Parallel Operation of Reverse Osmosis and Granular Activated Carbon for Drinking Water Treatment. Membranes 2021, 11, 33. [Google Scholar] [CrossRef] [PubMed]
  4. Dharupaneedi, S.P.; Sanna, K.N.; Mallikarjuna, N.; Kakarla, R.R.; Shyam, S.S.; Tejraj, M.A. Membrane-based separation of potential emerging pollutants. Sep. Purif. Technol. 2019, 210, 850–866. [Google Scholar] [CrossRef]
  5. Mayer, F.; Bhandari, R.; Gäth, S. Critical review on life cycle assessment of conventional and innovative waste-to-energy technologies. Sci. Total Environ. 2019, 672, 708–721. [Google Scholar] [CrossRef]
  6. Saravanan, A.; Deivayanai, V.C.; Senthil Kumar, P.; Rangasamy, G.; Hemavathy, R.V.; Harshana, T.; Gayathri, N.; Alagumalai, K. A detailed review on advanced oxidation process in treatment of wastewater: Mechanism, challenges and future outlook. Chemosphere 2022, 308, 136524. [Google Scholar] [CrossRef]
  7. Yu, H.Y.; Jenn, F.S.; Yujen, S.; Jianmin, W.; Po, Y.W.; Chin, P.H. Hazardous wastes treatment technologies. Water Environ. Res. 2020, 92, 1833–1860. [Google Scholar] [CrossRef]
  8. Lyngsie, G.; Krumina, L.; Tunlid, A.; Persson, P. Generation of hydroxyl radicals from reactions between a dimethoxyhydroquinone and iron oxide nanoparticles. Sci. Rep. 2018, 8, 10834. [Google Scholar] [CrossRef]
  9. Nidheesh, P.V.; Ganiyu, S.O.; Martínez-Huitle, C.A.; Mousset, E.; Olvera-Vargas, H.; Trellu, C.; Zhou, M.; Oturan, M.A. Recent advances in electro-Fenton process and its emerging applications. Crit. Rev. Environ. Sci. Technol. 2023, 53, 887–913. [Google Scholar] [CrossRef]
  10. Deng, F.; Jiang, J.; Sirés, I. State-of-the-art review and bibliometric analysis on electro-Fenton process. Carbon Lett. 2022, 33, 17–34. [Google Scholar] [CrossRef]
  11. Snyder, H. Literature review as a research methodology: An overview and guidelines. J. Bus. Res. 2019, 104, 333–339. [Google Scholar] [CrossRef]
  12. Grant, M.J.; Booth, A. A typology of reviews: An analysis of 14 review types and associated methodologies. Health Inf. Libr. J. 2009, 26, 91–108. [Google Scholar] [CrossRef] [PubMed]
  13. Hood, W.; Wilson, C.A. The Literature of Bibliometrics, Scientometrics, and Informetrics. Scientometrics 2001, 52, 291–314. [Google Scholar] [CrossRef]
  14. Jiang, Y.; Ran, J.; Mao, K.; Yang, X.; Zhong, L.; Yang, C.; Feng, X.; Zhang, H. Recent progress in Fenton/Fenton-like reactions for the removal of antibiotics in aqueous environments. Ecotoxicol. Environ. Saf. 2022, 236, 113464. [Google Scholar] [CrossRef] [PubMed]
  15. Wang, J.; Zhuan, R. Degradation of antibiotics by advanced oxidation processes: An overview. Sci. Total Environ. 2020, 701, 135023. [Google Scholar] [CrossRef]
  16. Nabout, J.C.; Parreira, M.R.; Teresa, F.B.; Carneiro, F.M.; Cunha, H.F.; Ondei, L.S.; Salomão Caramori, S.; Soares, T.N. Publish (in group) or perish (alone): The trend from single to multi-authorship in biological papers. Scientometrics 2015, 102, 357–364. [Google Scholar] [CrossRef]
  17. Marcionilio, S.M.L.O.; Alves, M.T.R.; Borges, P.P.; Machado, K.B.; Araújo, C.S.T.; Da Cunha, H.F.; Nabout, J.C. The state of global scientific literature on chlorophyll-A. Biosci. J. 2015, 31, 941–950. [Google Scholar] [CrossRef]
  18. Brillas, E. A review on the photoelectro-Fenton process as efficient electrochemical advanced oxidation for wastewater remediation. Treatment with UV light, sunlight, and coupling with conventional and other photo-assisted advanced technologies. Chemosphere 2020, 250, 126198. [Google Scholar] [CrossRef] [PubMed]
  19. Diaw, P.A.; Oturan, N.; Gaye Seye, M.D. Removal of the herbicide monolinuron from waters by the electro-Fenton treatment. J. Electroanal. Chem. 2020, 864, 114087. [Google Scholar] [CrossRef]
  20. Ramos, M.D.N.; Santana, C.S.; Velloso, C.C.V.; da Silva, A.H.M.; Magalhães, F.M.; Aguiar, A. A review on the treatment of textile industry effluents through Fenton processes. Process Saf. Environ. Prot. 2021, 155, 366–386. [Google Scholar] [CrossRef]
  21. Bu, J.; Liu, H.; Lin, C. Fenton’s reagent-enhanced supercritical water oxidation of wastewater released from 3-hydroxypyridine production. RSC Adv. 2019, 9, 29317–29326. [Google Scholar] [CrossRef] [PubMed]
  22. Varank, G.; Yazici, G.S.; Demir, A. A comparative study of electrocoagulation and electro-Fenton for food industry wastewater treatment: Multiple response optimization and cost analysis. Sep. Sci. Technol. 2018, 53, 2727–2740. [Google Scholar] [CrossRef]
  23. Ibarra-Taquez, H.N.; Dobrosz-Gómez, I.; Gómez, M.-Á. Optimización Multiobjetivo del Proceso Fenton en el Tratamiento de Aguas Residuales provenientes de la Producción de Café Soluble. Inf. Tecnol. 2018, 29, 111–122. [Google Scholar] [CrossRef]
  24. Ajmi, K.; Vismara, E.; Manai, I.; Haddad, M.; Hamdi, M.; Bouallagui, H. Polyvinyl acetate processing wastewater treatment using combined Fenton’s reagent and fungal consortium: Application of central composite design for conditions optimization. J. Hazard. Mater. 2018, 358, 243–255. [Google Scholar] [CrossRef] [PubMed]
  25. Dwivedi, K.; Morone, A.; Chakrabarti, T.; Pandey, R.A. Evaluation and optimization of Fenton pretreatment integrated with granulated activated carbon (GAC) filtration for carbamazepine removal from complex wastewater of pharmaceutical industry. J. Environ. Chem. Eng. 2018, 6, 3681–3689. [Google Scholar] [CrossRef]
  26. Abedinzadeh, N.; Shariat, M.; Monavari, S.M.; Pendashteh, A. Evaluation of color and COD removal by Fenton from biologically (SBR) pre-treated pulp and paper wastewater. Process Saf. Environ. Prot. 2018, 116, 82–91. [Google Scholar] [CrossRef]
  27. Cheng, Y.; Chen, Y.; Lu, J.; Nie, J.; Liu, Y. Fenton treatment of bio-treated fermentation-based pharmaceutical wastewater: Removal and conversion of organic pollutants as well as estimation of operational costs. Environ. Sci. Pollut. Res. 2018, 25, 12083–12095. [Google Scholar] [CrossRef]
  28. Gilpavas, E.; Arbeláez-Castaño, P.E.; Medina-Arroyave, J.D.; Gómez-Atehortua, C.M. Tratamiento de aguas residuales de la industria textil mediante coagulación química acoplada a procesos fenton intensificados con ultrasonido de baja frecuencia. Rev. Int. Contam. Ambient. 2018, 34, 157–167. [Google Scholar] [CrossRef]
  29. Ben Ayed, S.; Azam, M.; Al-Resayes, S.I.; Ayari, F.; Rizzo, L. Cationic Dye Degradation and Real Textile Wastewater Treatment by Heterogeneous Photo-Fenton, Using a Novel Natural Catalyst. Catalysts 2021, 11, 1358. [Google Scholar] [CrossRef]
  30. Oñate, J.; Arenas, A.; Ruiz, A.; Rivera, K.; Pelaez, C. Evaluation of Mutagenic and Genotoxic Activity in Vinasses Subjected to Different Treatments. Water Air Soil Pollut. 2015, 226, 144. [Google Scholar] [CrossRef]
  31. Ghosh, K.; Ghosh, J.; Giri, P.K. Accordion-like multilayered two-dimensional Ti3C2TxMXenes for catalytic elimination of organic dyes from wastewater via the Fenton reaction. ACS Appl. Nano Mater. 2022, 5, 16451–16461. [Google Scholar] [CrossRef]
  32. Yan, X.; Li, H.; Feng, J.; Hou, B.; Yan, W.; Zhou, M. Activated Carbon Assisted Fenton-like Treatment of Wastewater Containing Acid Red G. Catalysts 2022, 12, 1358. [Google Scholar] [CrossRef]
  33. Miao, S.; Gao, H.; Xia, H.; Mao, X.; Zhang, L.; Shi, M.; Zhang, Y. Accelerated Fenton degradation of azo dye wastewater via a novel Z-scheme CoFeN-g-C3N4 heterojunction photocatalyst with excellent charge transfer under visible light irradiation. Dalton Trans. 2022, 51, 17192–17202. [Google Scholar] [CrossRef]
  34. Sanabria, P.; Scunderlick, D.; Wilde, M.L.; Lüdtke, D.S.; Sirtori, C. Solar photo-Fenton treatment of the anti-cancer drug anastrozole in different aqueous matrices at near-neutral pH: Transformation products identification, pathways proposal, and in silico (Q)SAR risk assessment. Sci. Total Environ. 2021, 754, 142300. [Google Scholar] [CrossRef] [PubMed]
  35. Shen, X.; Cai, Z.; Hu, J.; Sun, B. Highly Efficient Microwave-Assisted Fenton Degradation of Toluene Nitration Wastewater over Microwave-Responsive Catalyst of Fe3O4−BiOCl. Chem. Sel. 2022, 7, e202200804. [Google Scholar] [CrossRef]
  36. Lin, Z.; Zhang, C.; Su, P.; Lu, W.; Zhang, Z.; Wang, X.; Hu, W. Fenton Process for Treating Acrylic Manufacturing Wastewater: Parameter Optimization, Performance Evaluation, Degradation Mechanism. Water 2022, 14, 2913. [Google Scholar] [CrossRef]
  37. Yazdanbakhsh, A.R.; Mohammadi, A.S.; Alinejad, A.A.; Hassani, G.; Golmohammadi, S.; Mohseni, S.M.; Sardar, M.; Sarsangi, V. Reduction of non-Betalactam Antibiotics COD by Combined Coagulation and Advanced Oxidation Processes. Water Environ. Res. 2016, 88, 2121–2131. [Google Scholar] [CrossRef]
  38. Nisapai, W.; Paikamnam, A.; Sriprom, P.; Neramittagapong, S.; Theerakulpisut, S.; Lin, C.; Neramittagapong, A. Degradation of Penicillin G contaminant in synthesized wastewater by Fenton-like reaction. Eng. Appl. Sci. Res. 2022, 49, 622–629. [Google Scholar]
  39. Yang, K.; Liu, M.; Weng, X.; Owens, G.; Chen, Z. Fenton-like oxidation for the simultaneous removal of estrone and β-estradiol from wastewater using biosynthesized silver nanoparticles. Sep. Purif. Technol. 2022, 285, 120304. [Google Scholar] [CrossRef]
  40. Li, H.; Liu, X.; Chen, X.; Chen, Y.; Li, Y.; Motkuri, R.K.; Dai, Z.; Kumar, A.; Fang, T.; Shen, J. Novel catalysts with multivalence copper for organic pollutants removal from wastewater with excellent selectivity and stability in Fenton-like process under neutral pH conditions. Water Environ. Res. 2022, 94, e10816. [Google Scholar] [CrossRef]
  41. Hassan, A.A.; Gheni, S.A.; Ahmed, S.M.R.; Abdullah, G.H.; Harvey, A. Aromatic free Fenton process for rapid removal of phenol from refinery wastewater in an oscillatory baffled reactor. Arab. J. Chem. 2022, 15, 103635. [Google Scholar] [CrossRef]
  42. Berberidou, C.; Kokkinos, P.; Poulios, I.; Mantzavinos, D. Homogeneous photo-Fenton degradation and mineralization of model and simulated pesticide wastewaters in lab- and pilot-scale reactors. Catalysts 2022, 12, 1512. [Google Scholar] [CrossRef]
  43. Lin, S.; Lu, Y.; Ye, B.; Zeng, C.; Liu, G.; Li, J.; Luo, H.; Zhang, R. Pesticide wastewater treatment using the combination of the microbial electrolysis desalination and chemical-production cell and Fenton process. Front. Environ. Sci. Eng. 2019, 14, 12. [Google Scholar] [CrossRef]
  44. Liu, W.; Yu, Y. A novel strategy for treating chromium complex wastewater: The combination of a Fenton-like reaction and adsorption using cobalt/iron-layered double hydroxide as catalyst and adsorbent. J. Clean. Prod. 2022, 370, 133337. [Google Scholar] [CrossRef]
  45. Arita, S.; Agustina, T.E.; Ilmi, N.; Pranajaya, V.D.W.; Gayatri, R. Treatment of laboratory wastewater by using Fenton reagent and combination of coagulation-adsorption as pretreatment. J. Ecol. Eng. 2022, 23, 8210–8220. [Google Scholar] [CrossRef]
  46. Cüce, H.; Aydin Temel, F. Classical-Fenton and photo-Fenton oxidation of wastewater arising from cosmetic automobile care products. Environ. Prog. Sustain. Energy 2021, 40, e13701. [Google Scholar] [CrossRef]
  47. Riadi, L.; Tanuwijaya, A.D.; Je, R.R.; Altway, A. Fenton’s Oxidation of Personal Care Product (PCP) Wastewater: A Kinetic Study and the Effects of System Parameters. Int. J. Technol. Manag. 2021, 12, 298–308. [Google Scholar] [CrossRef]
  48. Ferreira, L.C.; Salmerón, I.; Peres, J.A.; Tavares, P.B.; Lucas, M.S.; Malato, S. Advanced Oxidation Processes as sustainable technologies for the reduction of elderberry agro-industrial water impact. Water Resour. Ind. 2020, 24, 100137. [Google Scholar] [CrossRef]
  49. Pereira, M.; Oliveira, L.; Murad, E. Iron oxide catalysts: Fenton and Fenton-like reactions—A review. Clay Miner. 2012, 47, 285–302. [Google Scholar] [CrossRef]
  50. Xu, Y.; Guo, X.; Zha, F.; Tang, X.; Tian, H. Efficient photocatalytic removal of orange II by a Mn3O4-FeS2/Fe2O3 heterogeneous catalyst. J. Environ. Manag. 2020, 253, 109695. [Google Scholar] [CrossRef]
  51. Kumar, S.; Alka; Tarun; Saxena, J.; Bansal, C.; Kumari, P. Visible light-assisted photodegradation by silver tungstate-modified magnetite nanocomposite material for enhanced mineralization of organic water contaminants. Appl. Nanosci. 2020, 10, 1555–1569. [Google Scholar] [CrossRef]
  52. Fu, D.M.; Messele, S.A.; Fortuny, A.; Stuber, F.; Fabregat, A.; Font, J.; Bengoa, C. Efficient elimination of tyrosol in a zero valent iron-EDTA system at mild conditions. Chem. Eng. J. 2015, 260, 199–208. [Google Scholar] [CrossRef]
  53. Sirés, I.; Brillas, E. Upgrading and expanding the electro-Fenton and related processes. Curr. Opin. Electrochem. 2021, 27, 100686. [Google Scholar] [CrossRef]
  54. Chen, Y.; Cheng, Y.; Guan, X. A Rapid Fenton treatment of bio-treated dyeing and finishing wastewater at second-scale intervals: Kinetics by stopped-flow technique and application in a full-scale plant. Relat. Cient. 2019, 9, 9689. [Google Scholar] [CrossRef]
  55. Fischbacher, A.; Sonntag, C.; von Schmidt, T.C. Hydroxyl radical yields in the Fenton process under various pH, ligand concentrations and hydrogen peroxide/Fe(II) ratios. Chemosphere 2017, 182, 738–744. [Google Scholar] [CrossRef] [PubMed]
  56. Peng, L.; Duan, X.; Shang, Y.; Gao, B.; Xu, X. Engineered carbon supported single iron atom sites and iron clusters from Fe-rich Enteromorpha for Fenton-like reactions via nonradical pathways. Appl. Catal. B 2021, 287, 119963. [Google Scholar] [CrossRef]
  57. Ebrahiem, E.E.; Al-Maghrabi, M.N.; Mobarki, A.R. Removal of organic pollutants from industrial wastewater by applying photo-Fenton oxidation technology. Arab. J. Chem. 2017, 10, S1674–S1679. [Google Scholar] [CrossRef]
  58. Lisiée, M.; Gonçalves, P.; Eugênia, O.F.; Ostroski, I.C. Cosmetic wastewater primary treatment by fenton process and final polishing adsorption. Rev. Eletrôn. Gest. Educ. Tecnol. Ambient. 2020, 24, e13. [Google Scholar] [CrossRef]
  59. Li, S.; Gao, M.; Dong, H.; Jiang, Y.; Liang, W.; Jiang, J.; Ho, S.-H.; Li, F. Deciphering the fate of antibiotic resistance genes in norfloxacin wastewater treated by a bio-electro-Fenton system. Bioresour. Technol. 2022, 364, 128110. [Google Scholar] [CrossRef]
  60. Ribeiro, J.P.; Gomes, H.G.M.F.; Sarinho, L.; Marques, C.C.; Nunes, M.I. Synergies of metallic catalysts in the Fenton and photo-Fenton processes applied to the treatment of pulp bleaching wastewater. Chem. Eng. Process. Process Intensif. 2022, 181, 109159. [Google Scholar] [CrossRef]
  61. Ribeiro, J.P.; Sarinho, L.; Neves, M.C.; Nunes, M.I. Valorization of residual iron dust as Fenton catalyst for pulp and paper wastewater treatment. Environ. Pollut. 2022, 310, 119850. [Google Scholar] [CrossRef] [PubMed]
  62. Deng, F.; Olvera-Vargas, H.; Zhou, M.; Qiu, S.; Sirés, I.; Brillas, E. Critical Review on the Mechanisms of Fe2+ Regeneration in the Electro-Fenton Process: Fundamentals and Boosting Strategies. Chem. Rev. 2023, 123, 4635–4662. [Google Scholar] [CrossRef] [PubMed]
  63. Anastasiou, N.; Monou, M.; Mantzavinos, D.; Kassinos, D. Monitoring of the quality of winery influents/effluents and polishing of partially treated winery flows by homogeneous Fe(II) photo-oxidation. Desalination 2009, 248, 836–842. [Google Scholar] [CrossRef]
  64. De Torres-Socías, E.; Prieto-Rodríguez, L.; Zapata, A.; Fernández-Calderero, I.; Oller, I.; Malato, S. Detailed treatment line for a specific landfill leachate remediation. Brief economic assessment. Chem. Eng. J. 2015, 261, 60–66. [Google Scholar] [CrossRef]
  65. Hosseinzadeh, A.; Najafpoor, A.A.; Navaei, A.A.; Zhou, J.L.; Altaee, A.; Ramezanian, N.; Dehghan, A.; Bao, T.; Yazdani, M. Improving Formaldehyde Removal from Water and Wastewater by Fenton, Photo-Fenton and Ozonation/Fenton Processes through Optimization and Modeling. Water 2021, 13, 2754. [Google Scholar] [CrossRef]
  66. Hu, Y.; Yu, F.; Bai, Z.; Wang, Y.; Zhang, H.; Gao, X.; Wang, Y.; Li, X. Preparation of Fe-loaded needle coke particle electrodes and utilisation in three-dimensional electro-Fenton oxidation of coking wastewater. Chemosphere 2022, 308, 136544. [Google Scholar] [CrossRef] [PubMed]
  67. Afolabi, O.A.; Adekalu, K.O.; Okunade, D.A. Electro-Fenton treatment process for brewery wastewater: Effects of oxidant concentration and reaction time on BOD and COD removal efficiency. J. Eng. App. Sci. 2022, 69, 42. [Google Scholar] [CrossRef]
  68. Behrouzeh, M.; Mehdi Parivazh, M.; Danesh, E.; Javad Dianat, M.; Abbasi, M.; Osfouri, S.; Rostami, A.; Sillanpää, M.; Dibaj, M.; Akrami, M. Application of photo-Fenton, electro-Fenton, and photo-electro-Fenton processes for the treatment of DMSO and DMAC wastewaters. Arab. J. Chem. 2022, 15, 104229. [Google Scholar] [CrossRef]
  69. Martínez-Huitle, C.A.; Rodrigo, M.A.; Sirés, I.; Scialdone, O. A critical review on latest innovations and future challenges of electrochemical technology for the abatement of organics in water. Appl. Catal. B Environ. 2023, 328, 122430. [Google Scholar] [CrossRef]
  70. Perry, S.C.; Pangotra, D.; Vieira, L.; Csepei, L.-I.; Sieber, V.; Wang, L.; Ponce de León, C.; Walsh, F.C. Electrochemical synthesis of hydrogen peroxide from water and oxygen. Nat. Rev. Chem. 2019, 3, 442–458. [Google Scholar] [CrossRef]
  71. Zhou, W.; Meng, X.; Gao, J.; Alshawabkeh, A.N. Hydrogen peroxide generation from O2 electroreduction for environmental remediation: A state-of-the-art review. Chemosphere 2019, 225, 588–607. [Google Scholar] [CrossRef]
  72. Yang, W.; Zhou, M.; Oturan, N.; Li, Y.; Oturan, M.A. Electrocatalytic destruction of pharmaceutical imatinib by electro-Fenton process with graphene-based cathode. Electrochim. Acta 2019, 305, 285–294. [Google Scholar] [CrossRef]
  73. Coria, G.; Pérez, T.; Sirés, I.; Brillas, E.; Nava, J.L. Abatement of the antibiotic levofloxacin in a solar photoelectro-Fenton flow plant: Modeling the dissolved organic carbon concentration-time relationship. Chemosphere 2018, 198, 174–181. [Google Scholar] [CrossRef]
  74. Zhao, Q.; An, J.; Wang, S.; Qiao, Y.; Liao, C.; Wang, C.; Wang, X.; Li, N. Superhydrophobic air-breathing cathode for efficient hydrogen peroxide generation through two-electron pathway oxygen reduction reaction. ACS Appl. Mater. Interfaces 2019, 11, 35410–35419. [Google Scholar] [CrossRef]
  75. Xu, W.; Lu, Z.; Sun, X.; Jiang, L.; Duan, X. Superwetting electrodes for gas-involving electrocatalysis. Acc. Chem. Res. 2018, 51, 1590–1598. [Google Scholar] [CrossRef]
  76. Zhang, H.; Zhao, Y.; Li, A.Y.; Li, G.; Li, J.; Zhang, F. Janus electrode of asymmetric wettability for H2O2 production with highly efficient O2 utilization. ACS Appl. Mater. Interfaces 2020, 3, 705–714. [Google Scholar] [CrossRef]
  77. Zhang, H.; Li, Y.; Zhao, Y.; Li, G.; Zhang, F. Carbon black oxidized by air calcination for enhanced H2O2 generation and effective organics degradation. ACS Appl. Mater. Interfaces 2019, 11, 27846–27853. [Google Scholar] [CrossRef]
  78. Pérez, J.F.; Llanos, J.; Sáez, C.; López, C.; Cañizares, P.; Rodrigo, M.A. Electrochemical jet-cell for the in-situ generation of hydrogen peroxide. Electrochem. Comm. 2016, 71, 65–68. [Google Scholar] [CrossRef]
  79. Scialdone, O.; Galia, A.; Gattuso, C.; Sabatino, S.; Schiavo, B. Effect of air pressure on the electro-generation of H2O2 and the abatement of organic pollutants in water by electro-Fenton process. Electrochim. Acta 2015, 182, 775–780. [Google Scholar] [CrossRef]
  80. Hamdi, N.; Proietto, F.; Ben, A.H.; Galia, A.; Inguanta, R.; Ammar, S.; Gadri, A.; Scialdone, O. Effective Removal and mineralization of 8-hydroxyquinoline-5-sulfonic acid through a pressurized electro-Fenton-like process with Ni-Cu-Al layered double hydroxide. ChemElectroChem 2020, 7, 2457–2465. [Google Scholar] [CrossRef]
  81. Scialdone, O.; Galia, A.; Sabatino, S. Electro-generation of H2O2 and abatement of organic pollutant in water by an electro-Fenton process in a microfluidic reactor. Electrochem. Comm. 2013, 26, 45–47. [Google Scholar] [CrossRef]
  82. González, P.O.; Bisang, J.M. Electrochemical synthesis of hydrogen peroxide with a three-dimensional rotating cylinder electrode. J. Chem. Technol. Biotechnol. 2014, 89, 528–535. [Google Scholar] [CrossRef]
  83. Zhou, M.; Oturan, M.A.; Sirés, I. Electro-Fenton Process: New Trends and Scale-Up; Springer Nature: Singapore, 2018. [Google Scholar]
  84. Cornejo, O.M.; Sirés, I.; Nava, J.L. Electrosynthesis of hydrogen peroxide sustained by anodic oxygen evolution in a flow-through reactor. J. Electroanal. Chem. 2020, 873, 114419. [Google Scholar] [CrossRef]
  85. Guo, X.; Lin, S.; Gu, J.; Zhang, S.; Chem, Z.; Huang, S. Simultaneously achieving high activity and selectivity toward two-electron O2 electroreduction: The power of single-atom catalysts. ACS Catal. 2019, 9, 11042–11054. [Google Scholar] [CrossRef]
  86. Jung, E.; Shin, H.; Lee, B.-H.; Efremov, V.; Lee, S.; Lee, H.S.; Kim, J.; Antink, W.H.; Park, S.; Lee, K.-S.; et al. Atomic-level tuning of Co–N–C catalyst for high-performance electrochemical H2O2 production. Nat. Mater. 2020, 19, 436–442. [Google Scholar] [CrossRef]
  87. Pizzutilo, E.; Kasian, O.; Choi, C.H.; Cherevko, S.; Hutchings, G.J.; Mayrhofer, K.J.J.; Freakley, S.J. Electrocatalytic synthesis of hydrogen peroxide on Au-Pd nanoparticles: From fundamentals to continuous production. Chem. Phys. Lett. 2017, 683, 436–442. [Google Scholar] [CrossRef]
  88. Verdaguer-Casadevall, A.; Deiana, D.; Karamad, M.; Siahrostami, S.; Malacrida, P.; Hansen, T.W.; Rossmeisl, J.; Chorkendorff, I.; Stephens, I.E.L. Trends in the electrochemical synthesis of H2O2: Enhancing activity and selectivity by electrocatalytic site engineering. Nano Lett. 2014, 14, 1603–1608. [Google Scholar] [CrossRef]
  89. Colic, V.; Yang, S.; Révay, Z.; Stephens, I.E.L.; Chorkendorff, I. Carbon catalysts for electrochemical hydrogen peroxide production in acidic media. Electrochim. Acta 2018, 272, 192–202. [Google Scholar] [CrossRef]
  90. Chen, S.; Chen, Z.; Siahrostami, S.; Kim, T.R.; Nordlund, D.; Sokaras, D.; Nowak, S.; To, J.W.F.; Higgins, D.; Sinclair, R.; et al. Defective carbon-based materials for the electrochemical synthesis of hydrogen peroxide. ACS Sustain. Chem. Eng. 2018, 6, 311–317. [Google Scholar] [CrossRef]
  91. Barhoumi, N.; Olvera-Vargas, H.; Oturan, N.; Huguenot, D.; Gadri, A.; Ammar, S.; Brillas, E.; Oturan, M.A. Kinetics of oxidative degradation/mineralization pathways of the antibiotic tetracycline by the novel heterogeneous electro-Fenton process with solid catalyst chalcopyrite. Appl. Catal. B Environ. 2017, 209, 637–647. [Google Scholar] [CrossRef]
  92. Lanzalaco, S.; Sirés, I.; Sabatino, M.A.; Dispenza, C.; Scialdone, O.; Galia, A. Synthesis of polymer nanogels by electro-Fenton process: Investigation of the effect of main operation parameters. Electrochim. Acta 2017, 246, 812–822. [Google Scholar] [CrossRef]
  93. Liu, T.; Wang, K.; Song, S.; Brouzgou, A.; Tsiakaras, P.; Wang, Y. New electro-Fenton gas diffusion cathode based on nitrogen-doped graphene@carbon nanotube composite materials. Electrochim. Acta 2016, 194, 228–238. [Google Scholar] [CrossRef]
  94. Roth, H.; Gendel, Y.; Buzatu, P.; David, O.; Wessling, M. Tubular carbon nanotube-based gas diffusion electrode removes persistent organic pollutants by a cyclic adsorption electro-Fenton process. J. Hazard. Mater. 2016, 307, 1–6. [Google Scholar] [CrossRef] [PubMed]
  95. Mousset, E.; Ko, Z.T.; Syafik, M.; Wang, Z.; Lefebvre, O. Electro-catalytic activity enhancement of a graphene ink-coated carbon cloth cathode for oxidative treatment. Electrochim. Acta 2016, 222, 1628–1641. [Google Scholar] [CrossRef]
  96. Li, L.; Hu, H.; Teng, X.; Yu, Y.; Zhu, Y.; Su, X. Electrogeneration of H2O2 using a porous hydrophobic acetylene black cathode for electro-Fenton process. Chem. Eng. Process. Process Intensif. 2018, 133, 34–39. [Google Scholar] [CrossRef]
  97. Zhao, K.; Su, Y.; Quan, X.; Liu, Y.; Chen, S.; Yu, H. Enhanced H2O2 production by selective electrochemical reduction of O2 fluorine-doped hierarchically porous carbon. J. Catal. 2018, 357, 118–126. [Google Scholar] [CrossRef]
  98. Yu, F.; Wang, Y.; Ma, H. Enhancing the yield of H2O2 from oxygen reduction reaction performance by hierarchically porous carbon modified active carbon fiber as an effective cathode used in electro-Fenton. J. Electroanal. Chem. 2019, 838, 57–65. [Google Scholar] [CrossRef]
  99. Wang, Y.; Zhou, W.; Gao, J.; Ding, Y.; Kou, K. Oxidative modification of graphite felts for efficient H2O2 electrogeneration: Enhancement mechanism and long-term stability. J. Electroanal. Chem. 2019, 833, 258–268. [Google Scholar] [CrossRef]
  100. Zhang, H.; Li, Y.; Li, G.; Zhang, F. Scaling up floating air cathodes for energy-efficient H2O2 generation and electrochemical advanced oxidation processes. Electrochim. Acta 2019, 299, 273–280. [Google Scholar] [CrossRef]
  101. Perazzolo, V.; Durante, C.; Gennaro, A. Nitrogen and sulfur doped mesoporous carbon cathodes for water treatment. J. Electroanal. Chem. 2016, 782, 264–269. [Google Scholar] [CrossRef]
  102. Xia, Y.; Shang, H.; Zhang, Q.; Zhou, Y.; Hu, X. Electrogeneration of hydrogen peroxide using phosphorus-doped carbon nanotubes gas diffusion electrodes and its application in electro-Fenton. J. Electroanal. Chem. 2019, 840, 400–408. [Google Scholar] [CrossRef]
  103. Zhu, Y.; Qiu, S.; Deng, F.; Ma, F.; Zheng, Y. Degradation of sulfathiazole by electro-Fenton using a nitrogen-doped cathode and a BDD anode: Insight into the H2O2 generation and radical oxidation. Sci. Total Environ. 2020, 722, 137853. [Google Scholar] [CrossRef] [PubMed]
  104. Rocha, R.S.; Silva, F.L.; Valim, R.B.; Barros, W.R.P.; Steter, J.R.; Bertazzoli, R.; Lanza, M.R.V. Effect of Fe2+ on the degradation of the pesticide profenofos by electrogenerated H2O2. J. Electroanal. Chem. 2016, 783, 100–105. [Google Scholar] [CrossRef]
  105. Moreira, J.; Lima, V.B.; Goulart, L.A.; Lanza, M.R.V. Electrosynthesis of hydrogen peroxide using modified gas diffusion electrodes (MGDE) for environmental applications: Quinones and azo compounds employed as redox modifiers. Appl. Catal. B Environ. 2019, 248, 95–107. [Google Scholar] [CrossRef]
  106. Jirkovský, J.S.; Björling, A.; Ahlberg, E. Reduction of oxygen on dispersed nanocrystalline CoS2. J. Phys. Chem. C 2012, 116, 24436–24444. [Google Scholar] [CrossRef]
  107. Ye, Z.; Brillas, E.; Centellas, F.; Cabot, P.L.; Sirés, I. Electro-Fenton process at mild pH using Fe(III)-EDDS as soluble catalyst and carbon felt as cathode. Appl. Catal. B Environ. 2019, 257, 117907. [Google Scholar] [CrossRef]
  108. Alcaide, F.; Álvarez, G.; Guelfi, D.R.V.; Brillas, E.; Sirés, I. A stable CoSP/MWCNTs air-diffusion cathode for the photoelectro-Fenton degradation of organic pollutants at pre-pilot scale. Chem. Eng. J. 2020, 379, 122417. [Google Scholar] [CrossRef]
  109. Antonin, V.S.; Parreira, L.S.; Aveiro, L.R.; Silva, F.L.; Valim, R.B.; Hammer, P.; Lanza, M.R.V.; Santos, M.C. W@Au Nanostructures modifying carbon as materials for hydrogen peroxide electrogeneration. Electrochim. Acta 2017, 231, 713–720. [Google Scholar] [CrossRef]
  110. Sajjadi, S.; Hasanzadeh, A.; Khataee, A. Two-electron oxygen reduction on NiFe alloy enclosed carbonic nanolayers derived from NiFe-metal-organic frameworks. J. Electroanal. Chem. 2019, 840, 449–455. [Google Scholar] [CrossRef]
  111. Xu, A.; He, B.; Yu, H.; Han, W.; Li, J.; Shen, J.; Sun, X.; Wang, L. A facile solution to mature cathode modified by hydrophobic dimethyl silicon oil (DMS) layer for electro-Fenton processes: Water proof and enhanced oxygen transport. Electrochim. Acta 2019, 308, 158–166. [Google Scholar] [CrossRef]
  112. Zhang, Q.; Zhou, M.; Ren, G.; Li, Y.; Li, Y.; Du, X. Highly efficient electrosynthesis of hydrogen peroxide on a super-hydrophobic three-phase interface by natural air diffusion. Nat. Commun. 2020, 11, 1731. [Google Scholar] [CrossRef]
  113. Borghei, M.; Lehtonen, J.; Liu, L.; Rojas, O.J. Advanced biomass-derived electrocatalysts for the oxygen reduction reaction. Adv. Mater. 2018, 30, 1703691. [Google Scholar] [CrossRef] [PubMed]
  114. Liao, M.-J.; Wang, Y.-L.; Li, S.-S.; Li, J.-F.; Chen, P. Electrocatalyst derived from abundant biomass and its excellent activity for in situ H2O2 production. ChemElectroChem 2019, 6, 4877–4884. [Google Scholar] [CrossRef]
  115. Liang, L.; Zhou, M.; Lu, X.; Su, P.; Sun, J. High-efficiency electro-generation of hydrogen peroxide from oxygen reduction by carbon xerogels derived from glucose. Electrochim. Acta 2019, 320, 134569. [Google Scholar] [CrossRef]
  116. Zhang, H.-X.; Yang, S.-C.; Wang, Y.-L.; Xi, J.-C.; Huang, J.-C.; Li, J.-F.; Chen, P.; Jia, R. Electrocatalyst derived from fungal hyphae and its excellent activity for electrochemical production of hydrogen peroxide. Electrochim. Acta 2019, 308, 74–82. [Google Scholar] [CrossRef]
  117. Daniel, G.; Zhang, Y.; Lanzalaco, S.; Brombin, F.; Kosmala, T.; Granozzi, G.; Wang, A.; Brillas, E.; Sirés, I.; Durante, C. Chitosan-derived nitrogen-doped carbon electrocatalyst for a sustainable upgrade of oxygen reduction to hydrogen peroxide in UV-assisted electro-Fenton water treatment. ACS Sustain. Chem. Eng. 2020, 8, 14425–14440. [Google Scholar] [CrossRef]
  118. Ganiyu, S.O.; Zhou, M.; Martínez-Huitle, C.A. Heterogeneous electro-Fenton and photoelectro-Fenton processes: A critical review of fundamental principles and application for water/wastewater treatment. Appl. Catal. B Environ. 2018, 235, 103–129. [Google Scholar] [CrossRef]
  119. Zhou, W.; Rajic, L.; Zhao, Y.; Gao, J.; Qin, Y.; Alshawabkeh, A. Rates of H2O2 electrogeneration by reduction of anodic O2 at RVC foam cathodes in batch and flow-through cells. Electrochim. Acta 2018, 277, 185–196. [Google Scholar] [CrossRef]
  120. Ye, Z.; Guelfi, D.R.V.; Álvarez, G.; Alcaide, F.; Brillas, E.; Sirés, I. Enhanced electrocatalytic production of H2O2 at Co-based air-diffusion cathodes for the photoelectro-Fenton treatment of bronopol. Appl. Catal. B Environ. 2019, 247, 191–199. [Google Scholar] [CrossRef]
  121. Ye, Z.; Brillas, E.; Centellas, F.; Cabot, P.L.; Sirés, I. Expanding the application of photoelectro-Fenton treatment to urban wastewater using the Fe(III)-EDDS complex. Water Res. 2020, 169, 115219. [Google Scholar] [CrossRef]
  122. Labiah, L.; Oturan, M.A.; Panizza, M.; Ben Hamadi, N.; Ammar, S. Complete removal of AHPS synthetic dye from water using new electro-Fenton oxidation catalyzed by natural pyrite as heterogeneous catalyst. J. Hazard. Mater. 2015, 297, 34–41. [Google Scholar] [CrossRef] [PubMed]
  123. Droguett, C.; Salazar, R.; Brillas, E.; Sirés, I.; Carlesi, C.; Marco, J.F.; Thiam, A. Treatment of antibiotic cephalexin by heterogeneous electrochemical Fenton-based processes using chalcopyrite as sustainable catalyst. Sci. Total Environ. 2020, 740, 140154. [Google Scholar] [CrossRef] [PubMed]
  124. Dos Santos, A.J.; Sirés, I.; Alves, A.P.M.; Martínez-Huitle, C.A.; Brillas, E. Vermiculite as heterogeneous catalyst in electrochemical Fenton-based processes: Application to the oxidation of Ponceau SS dye. Chemosphere 2020, 240, 124838. [Google Scholar] [CrossRef] [PubMed]
  125. Sklari, S.D.; Plakas, K.V.; Petsi, P.N.; Zaspalis, V.T.; Karabelas, A.J. Toward the development of a novel electro-Fenton System for eliminating toxic organic substances from water. Part 2. Preparation, characterization, and evaluation of iron-impregnated carbon felts as cathodic electrodes. Ind. Eng. Chem. Res. 2015, 54, 2059–2073. [Google Scholar] [CrossRef]
  126. Fernández, D.; Robles, I.; Rodríguez-Valadez, F.J.; Godínez, L.A. Novel arrangement for an electro-Fenton reactor that does not require addition of iron, acid and a final neutralization stage. Towards the development of a cost-effective technology for the treatment of wastewater. Chemosphere 2018, 199, 251–255. [Google Scholar] [CrossRef] [PubMed]
  127. Rostamizadeh, M.; Jafarizad, A.; Gharibian, S. High efficient decolorization of Reactive Red 120 azo dye over reusable Fe-ZSM-5 nanocatalyst in electro-Fenton reaction. Separ. Purif. Technol. 2018, 192, 340–347. [Google Scholar] [CrossRef]
  128. Ye, Z.; Padilla, J.A.; Xuriguera, E.; Brillas, E.; Sirés, I. Magnetic MIL(Fe)-type MOF-derived N-doped nano-ZVI@C rods as heterogeneous catalyst for the electro-Fenton degradation of gemfibrozil in a complex aqueous matrix. Appl. Catal. B Environ. 2020, 266, 118604. [Google Scholar] [CrossRef]
  129. Ye, Z.; Schukraft, G.E.M.; L’Hermitte, A.; Xiong, Y.; Brillas, E.; Petit, C.; Sirés, I. Mechanism and stability of an Fe-based 2D MOF during the photoelectro-Fenton treatment of organic micro-pollutants under UVA and visible light irradiation. Water Res. 2020, 184, 115986. [Google Scholar] [CrossRef]
  130. Cheng, Z.; Li, L.; Liu, J. Industrial structure, technical progress and carbon intensity in China’s provinces. Renew. Sust. Energ. Rev. 2018, 81, 2935–2946. [Google Scholar] [CrossRef]
  131. Nan, X.; Lavrnić, S.; Toscano, A. Potential of constructed wetland treatment systems for agricultural wastewater reuse under the EU framework. J. Environ. Manag. 2020, 275, 111219. [Google Scholar] [CrossRef]
  132. Mei, N.S.; Wai, C.W.; Ahamad, R. Environmental Awareness and Behaviour Index for Malaysia. Procedia Soc. Behav. Sci. 2016, 222, 668–675. [Google Scholar] [CrossRef]
  133. Silva, M.M.; Luiz, G.P.; Duarte, S.; Higor, H.C. Práticas de gerenciamento de resíduos industriais no Brasil: Uma revisão da literatura. Rev. Bras. Eng. Prod. 2019, 5, 251–261. [Google Scholar]
  134. Maryam, B.; Büyükgüngör, H. Wastewater reclamation and reuse trends in Turkey: Opportunities and challenges. J. Water Process Eng. 2019, 30, 100501. [Google Scholar] [CrossRef]
  135. Domingues, E.; Fernandes, E.; Vaz, T.; Gomes, J.; Castro-Silva, S.; Martins, R.C.; Quinta-Ferreira, R.; Ferreira, L.M. Ion Exchange to Capture Iron after Real Effluent Treatment by Fenton’s Process. Water 2022, 14, 706. [Google Scholar] [CrossRef]
  136. Volesky, B. Detoxification of metal-bearing effluents: Biosorption for the next century. Hydrometallurgy 2001, 59, 203–216. [Google Scholar] [CrossRef]
  137. Martins, P.J.M.; Reis, P.M.; Martins, R.C.; Gando-Ferreira, L.M.; Quinta-Ferreira, R.M. Iron recovery from the Fenton’s treatment of winery effluent using an ion-exchange resin. J. Mol. Liq. 2017, 242, 505–511. [Google Scholar] [CrossRef]
  138. Hasani, K.; Peyghami, A.; Moharrami, A.; Vosoughi, M.; Dargahi, A. The efficacy of sono-electro-Fenton process for removal of cefixime antibiotic from aqueous solutions by response surface methodology (RSM) and evaluation of toxicity of effluent by microorganisms. Arab. J. Chem. 2020, 13, 6122–6139. [Google Scholar] [CrossRef]
  139. Garcia-Segura, S.; Bellotindos, L.M.; Huang, Y.-H.; Brillas, E.; Lu, M.-C. Fluidized-bed Fenton process as alternative wastewater treatment technology—A review. J. Taiwan Inst. Chem. 2016, 67, 211–225. [Google Scholar] [CrossRef]
  140. Lacson, C.F.Z.; de Luna, M.D.G.; Dong, C.; Garcia-Segura, S.; Lu, M.-C. Fluidized-bed Fenton treatment of imidacloprid: Optimization and degradation pathway. Sustain. Environ. Res. 2018, 28, 309–314. [Google Scholar] [CrossRef]
  141. Su, C.-C.; Pukdee-Asa, M.; Ratanatamskul, C.; Lu, M.-C. Effect of operating parameters on decolorization and COD removal of three reactive dyes by Fenton’s reagent using fluidized-bed reactor. Desalination 2011, 278, 211–218. [Google Scholar] [CrossRef]
  142. Arden, S.; Ma, X. Constructed wetlands for greywater recycle and reuse: A review. Sci. Total Environ. 2018, 630, 587–599. [Google Scholar] [CrossRef]
  143. Liu, T.; Xu, S.; Lu, S.; Qin, P.; Bi, B.; Ding, H.; Liu, Y.; Guo, X.; Liu, X. A review on removal of organophosphorus pesticides in constructed wetland: Performance, mechanism and influencing factors. Sci. Total Environ. 2019, 651, 2247–2268. [Google Scholar] [CrossRef] [PubMed]
  144. Ahmad, N.N.R.; Ang, W.L.; Teow, Y.H.; Mohammand, A.W.; Hilal, N. Nanofiltration membrane processes for water recyclinh, reuse and product recovery within various industries: A review. J. Water Process Eng. 2022, 45, 102478. [Google Scholar] [CrossRef]
  145. Ali, M.; Rousseau, D.P.L.; Ahmed, S. A full-scale comparison of two hybrid constructed wetlands treating domestic wastewater in Pakistan. J. Environ. Manag. 2018, 210, 349–358. [Google Scholar] [CrossRef] [PubMed]
  146. Gorito, A.M.; Ribeiro, A.R.; Gomes, C.R.; Almeida, C.M.R.; Silva, A.M.T. Constructed wetland microcosms for the removal of organic micropollutants from freshwater aquaculture effluents. Sci. Total Environ. 2018, 644, 1171–1180. [Google Scholar] [CrossRef] [PubMed]
  147. Hussain, Z.; Arslan, M.; Malik, M.H.; Mohsin, M.; Iqbal, S.; Afzal, M. Treatment of the textile industry effluent in a pilot-scale vertical flow constructed wetland system augmented with bacterial endophytes. Sci. Total Environ. 2018, 645, 966–973. [Google Scholar] [CrossRef]
  148. Maine, M.A.; Sanchez, G.C.; Hadad, H.R.; Caffaratti, S.E.; Pedro, M.C.; Mufarrege, M.M.; Di Luca, G.A. Hybrid constructed wetlands for the treatment of wastewater from a fertilizer manufacturing plant: Microcosms and field scale experiments. Sci. Total Environ. 2019, 650, 297–302. [Google Scholar] [CrossRef]
  149. Bang, W.H.; Jung, Y.; Park, J.W.; Lee, S.; Maeng, S.K. Effects of hydraulic loading rate and organic load on the performance of a pilot-scale hybrid VF-HF constructed wetland in treating secondary effluent. Chemosphere 2019, 218, 232–240. [Google Scholar] [CrossRef]
  150. Tang, X.Y.; Yang, Y.; McBride, M.B.; Tao, R.; Dai, Y.N.; Zhang, X.M. Removal of chlorpyrifos in recirculating vertical flow constructed wetlands with five wetland plant species. Chemosphere 2019, 216, 195–202. [Google Scholar] [CrossRef]
  151. Xing, L.; Kong, M.; Xie, X.; Sun, J.; Wei, D.; Li, A. Feasibility and safety of papermaking wastewarter in using as ecological water supplement after advanced treatment by fluidized-bed Fenton coupled with large-scale constructed wetland. Sci. Total Environ. 2020, 699, 134369. [Google Scholar] [CrossRef]
  152. Cai, Q.Q.; Lee, B.C.Y.; Ong, S.L.; Hu, J.Y. Fleuidized-bed Fenton techonologies for recalcitrant industrial wastewater treatment-Recent advances, chellenges and perspective. Water Res. 2021, 190, 116692. [Google Scholar] [CrossRef] [PubMed]
  153. Metin, S.; Çifçi, D.İ. Chemical industry wastewater treatment by coagulation combined with Fenton and photo-Fenton processes. J. Chem. Technol. Biotechnol. 2023, 98, 1158–1165. [Google Scholar] [CrossRef]
  154. Lin, R.; Li, Y.; Yong, T.; Cao, W.; Wu, J.; Shen, Y. Synergistic effects of oxidation, coagulation and adsorption in the integrated fenton-based process for wastewater treatment: A review. J. Environ. Manag. 2022, 306, 114460. [Google Scholar] [CrossRef]
  155. Sun, G.; Zhang, Y.; Gao, Y.; Han, X.; Yang, M. Removal of hard COD from biological effluent of coking wastewater using synchronized oxidation-adsorption technology: Performance, mechanism, and full-scale application. Water Res. 2020, 173, 115517. [Google Scholar] [CrossRef]
  156. Niaounakis, M.; Halvadakis, C.P. Olive processing waste management literature review and patent survey. Waste Manag. Ser. 2006, 5, 23–64. [Google Scholar]
  157. Paraskeva, P.; Diamadopoulos, E. Technologies for olive mill wastewater (OMW) treatment: A review. J. Chem. Technol. Biotechnol. 2006, 81, 475–1485. [Google Scholar] [CrossRef]
  158. Asfi, M.; Ouzounidou, G.; Panajiotidis, S.; Therios, I.; Moustakas, M. Toxicity effects of olive-mill wastewater on growth, photosynthesis and pollen morphology of spinach plants. Ecotoxicol. Environ. Saf. 2012, 80, 69–75. [Google Scholar] [CrossRef]
  159. Danellakis, D.; Ntaikou, I.; Kornaros, M.; Dailianis, S. Olive oil mill wastewater toxicity in the marine environment: Alterations of stress indices in tissues of mussel Mytilus galloprovincialis. Aquat. Toxicol. 2011, 101, 358–366. [Google Scholar] [CrossRef] [PubMed]
  160. Karaouzas, I.; Skoulikidis, N.T.; Giannakou, U.; Albanis, T.A. Spatial and temporal effects of olive mill wastewaters to stream macroinvertebrates and aquatic ecosystems status. Water Res. 2011, 45, 6334–6346. [Google Scholar] [CrossRef] [PubMed]
  161. Ntougias, S.; Gaitis, F.; Katsaris, P.; Skoulika, S.; Iliopoulos, N.; Zervakis, G.I. The effects of olives harvest period and production year on olive mill wastewater properties-evaluation of Pleurotus strains as bioindicators of the effluent’s toxicity. Chemosphere 2013, 92, 399–405. [Google Scholar] [CrossRef]
  162. Ochando-Pulido, J.M.; Victor-Ortega, M.D.; Hodaifa, G.; Martinez-Ferez, A. Physicochemical analysis and adequation of olive oil mill wastewater after advanced oxidation process for reclamation by pressure-driven membrane technology. Sci. Total Environ. 2015, 503–504, 113–121. [Google Scholar] [CrossRef] [PubMed]
  163. De Caprariis, B.; Di Rita, M.; Stoller, M.; Verdone, N.; Chianese, A. Reaction-precipitation by a spinning disc reactor: Influence of hydrodynamics on nanoparticles production. Chem. Eng. Sci. 2012, 76, 73–80. [Google Scholar] [CrossRef]
  164. Martínez Nieto, L.; Hodaifa, G.; Rodríguez Vives, S.; Giménez Casares, J.A.; Ochando, J. Flocculation–sedimentation combined with chemical oxidation process. Clean Soil Air Water 2011, 39, 949–955. [Google Scholar] [CrossRef]
  165. Sacco, O.; Stoller, M.; Vaiano, V.; Ciambelli, P.; Chianese, A.; Sannino, D. Photocatalytic degradation of organic dyes under visible light on n-doped photocatalysts. Int. J. Photoenergy 2012, 2012, 626759. [Google Scholar] [CrossRef]
  166. Martínez Nieto, L.; Hodaifa, G.; Rodríguez, V.S.; Giménez, C.J.A.; Ochando, J. Degradation of organic matter in olive oil mill wastewater through homogeneous Fenton-like reaction. Chem. Eng. J. 2011, 173, 503–510. [Google Scholar] [CrossRef]
  167. Martínez Nieto, L.; Alami, S.B.D.; Hodaifa, G.; Faour, C.; Rodríguez, S.; Gimézez, J.A.; Ochando, J. Adsorption of iron on crude olive stones. Ind. Crop. Prod. 2010, 32, 467–471. [Google Scholar] [CrossRef]
  168. Hodaifa, G.; Ochando-Pulido, J.M.; Rodriguez-Vives, S.; Martinez-Ferez, A. Optimization of con- tinuous reactor at pilot scale for olive-oil mill wastewater treatment by Fenton-like process. Chem. Eng. J. 2013, 220, 117–124. [Google Scholar] [CrossRef]
  169. Hodaifa, G.; Ochando-Pulido, J.M.; Ben-Driss-Alami, S.; Rodriguez-Vives, S.; Martinez-Ferez, A. Kinetic and thermodynamic parameters of iron adsorption onto olive stones. Ind. Crop. Prod. 2013, 49, 526–534. [Google Scholar] [CrossRef]
  170. Sancho Cierva, J. Water Quality for Irrigation Use; Giner, J.F., Ed.; Universida Politécnica de Valencia-Generalitat Valenciana—Phytoma: Valencia, Spain, 2000. [Google Scholar]
  171. Iaquinta, M.; Stoller, M.; Merli, C. Optimization of a nanofiltration membrane for tomato industry wastewater treatment. Desalination 2009, 245, 314–320. [Google Scholar] [CrossRef]
  172. Stoller, M.; Chianese, A. Technical optimization of a batch olive wash wastewater treatment membrane plant. Desalination 2006, 200, 734–736. [Google Scholar] [CrossRef]
  173. Stoller, M.; Chianese, A. Optimization of membrane batch processes by means of the critical flux theory. Desalination 2006, 191, 62–70. [Google Scholar] [CrossRef]
  174. Stoller, M.; Chianese, A. Influence of the adopted pretreatment process on the critical flux value of batch membrane processes. Ind. Eng. Chem. Res. 2007, 46, 2249–2253. [Google Scholar] [CrossRef]
  175. Stoller, M.; Bravi, M.; Chianese, A. Threshold flux measurements of a nanofiltration mem- brane module by critical flux data conversion. Desalination 2013, 315, 142–148. [Google Scholar] [CrossRef]
  176. Stoller, M.; De Caprariis, B.; Cicci, A.; Verdone, N.; Bravi, M.; Chianese, A. About proper mem- brane process design affected by fouling by means of the analysis of measured threshold flux data. Sep. Purif. Technol. 2013, 114, 83–89. [Google Scholar] [CrossRef]
  177. Stoller, M. On the effect of flocculation as pretreatment process and particle size distribution for membrane fouling reduction. Desalination 2009, 240, 209–217. [Google Scholar] [CrossRef]
  178. Stoller, M. Effective fouling inhibition by critical flux based optimization methods on a NF membrane module for olive mill wastewater treatment. Chem. Eng. J. 2011, 168, 1140–1148. [Google Scholar] [CrossRef]
  179. Pal, M.; Malhotra, M.; Madal, M.K.; Paine, T.K.; Pal, P. Recycling of wastewater from tannery industry through membrane-integrated hybrid treatment using a novel graphene oxide nanocomposite. J. Water Process. Eng. 2020, 36, 101324. [Google Scholar] [CrossRef]
  180. Víctor-Ortega, M.D.; Ochando-Pulido, J.M.; Hodaifa, G.; Martinez-Ferez, A. Final purification of synthetic olive oil mill wastewater treated by chemical oxidation using ion exchange: Study of operating parameters. Chem. Eng. Process. Process Intensif. 2014, 85, 241–247. [Google Scholar] [CrossRef]
  181. El-Hazek, A.N.; Wagdy, A.H.; Hassanain, A.M. Assessment of a Fenton Reaction in treating Greywater for reuse in Irrigation. Eng. Res. J. 2022, 51, 141–148. [Google Scholar]
  182. Ribeiro, M.C.M.; Starling, M.C.V.M.; Leão, M.M.D.; Amorim, C.C. Txtile wastewater reuse after additional treatment by Fenton’s reagent. Environ. Sci. Pollut. Res. 2016, 24, 6165–6175. [Google Scholar] [CrossRef]
Figure 1. Comparison of the number of articles, between WoS and Scopus databases, using the same search methodology. Temporal trend in number of articles published in WoS and Scopus database.
Figure 1. Comparison of the number of articles, between WoS and Scopus databases, using the same search methodology. Temporal trend in number of articles published in WoS and Scopus database.
Processes 11 02466 g001
Figure 2. Temporal trend in number of articles published between 1992 and 2022 in Scopus database, considering origin of effluents degraded by Fenton.
Figure 2. Temporal trend in number of articles published between 1992 and 2022 in Scopus database, considering origin of effluents degraded by Fenton.
Processes 11 02466 g002
Figure 3. Fenton reaction arrangements on 30-year time scale.
Figure 3. Fenton reaction arrangements on 30-year time scale.
Processes 11 02466 g003
Table 1. Temporal trend in number of articles published between 1992 and 2022, and Pearson correlation (r) in relation to number of articles published per year × type of compounds degraded by Fenton.
Table 1. Temporal trend in number of articles published between 1992 and 2022, and Pearson correlation (r) in relation to number of articles published per year × type of compounds degraded by Fenton.
VariableNo. of Articlesrp
No. of articles 12630.93<0.001
Dye 2270.85<0.001
Drugs 1140.81<0.001
Phenol 620.81<0.001
Organic compounds 2160.87<0.001
Hygiene and cleaning 110.420.02
Pesticide 220.59<0.001
Inorganic compounds/metals 400.82<0.001
Vinasse 50.320.09
Other compounds 4180.69<0.001
Table 2. List of 23 journals with the most studies on Fenton until 2022.
Table 2. List of 23 journals with the most studies on Fenton until 2022.
JournalTotal%
Desalination and Water Treatment685.384
Chinese Journal of Environmental Engineering675.305
Journal of Hazardous Materials665.226
Chemical Engineering Journal564.434
Chemosphere383.009
Journal of Environmental Chemical Engineering342.692
Water Science and Technology342.692
Environmental Science and Pollution Research272.138
Journal of Environmental Management251.979
Separation and Purification Technology251.979
Water Research231.821
Environmental Technology (United Kingdom)181.425
Shenyang Jianzhu Daxue Xuebao (Ziran Kexue Ban)/Journal of Shenyang Jianzhu University (Natural Science)171.346
Journal of Water Process Engineering161.267
RSC Advances141.108
Science of the Total Environment141.108
Chung-kuo Tsao Chih/China Pulp and Paper131.029
Journal of Cleaner Production131.029
Journal of Chemical Technology and Biotechnology120.950
Desalination110.871
International Journal of Environmental Science and Technology110.871
Asian Journal of Chemistry100.792
Processes100.792
Table 3. List of 10 countries that published the most studies on Fenton indexed in SCOPUS until 2022.
Table 3. List of 10 countries that published the most studies on Fenton indexed in SCOPUS until 2022.
CountryNo. of ArticlesPercentage
China53041.96358
Turkey957.52177
Iran866.80918
Spain624.90895
India524.11718
Portugal433.40459
Taiwan383.00871
Brazil292.29612
Poland272.13777
Italy181.42518
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Pimentel Prates, M.; de Oliveira Loures Marcionílio, S.M.; Borges Machado, K.; Medeiros de Araújo, D.; Martínez-Huitle, C.A.; Leão Arantes, A.L.; Ferreira da Silva Gadêlha, J.E. Fenton: A Systematic Review of Its Application in Wastewater Treatment. Processes 2023, 11, 2466. https://doi.org/10.3390/pr11082466

AMA Style

Pimentel Prates M, de Oliveira Loures Marcionílio SM, Borges Machado K, Medeiros de Araújo D, Martínez-Huitle CA, Leão Arantes AL, Ferreira da Silva Gadêlha JE. Fenton: A Systematic Review of Its Application in Wastewater Treatment. Processes. 2023; 11(8):2466. https://doi.org/10.3390/pr11082466

Chicago/Turabian Style

Pimentel Prates, Matheus, Suzana Maria de Oliveira Loures Marcionílio, Karine Borges Machado, Danyelle Medeiros de Araújo, Carlos A. Martínez-Huitle, Arizeu Luiz Leão Arantes, and José Eduardo Ferreira da Silva Gadêlha. 2023. "Fenton: A Systematic Review of Its Application in Wastewater Treatment" Processes 11, no. 8: 2466. https://doi.org/10.3390/pr11082466

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop