Next Article in Journal
Catalytic Processes for Water and Wastewater Treatment
Next Article in Special Issue
Supported Ni Single-Atom Catalysts: Synthesis, Structure, and Applications in Thermocatalytic Reactions
Previous Article in Journal
Pt/CeMnOx/Diatomite: A Highly Active Catalyst for the Oxidative Removal of Toluene and Ethyl Acetate
Previous Article in Special Issue
Oxygen-Deficient Engineering for Perovskite Oxides in the Application of AOPs: Regulation, Detection, and Reduction Mechanism
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

From Fenton and ORR 2e-Type Catalysts to Bifunctional Electrodes for Environmental Remediation Using the Electro-Fenton Process

by
Edgar Fajardo-Puerto
,
Abdelhakim Elmouwahidi
,
Esther Bailón-García
*,
Agustín Francisco Pérez-Cadenas
and
Francisco Carrasco-Marín
Materiales Polifuncionales Basados en Carbono (UGR-Carbon), Departamento de Química Inorgánica, Universidad de Granada, 18071 Granada, Spain
*
Author to whom correspondence should be addressed.
Catalysts 2023, 13(4), 674; https://doi.org/10.3390/catal13040674
Submission received: 6 March 2023 / Revised: 23 March 2023 / Accepted: 24 March 2023 / Published: 30 March 2023
(This article belongs to the Special Issue Exclusive Review Papers in Catalytic Materials)

Abstract

:
Currently, the presence of emerging contaminants in water sources has raised concerns worldwide due to low rates of mineralization, and in some cases, zero levels of degradation through conventional treatment methods. For these reasons, researchers in the field are focused on the use of advanced oxidation processes (AOPs) as a powerful tool for the degradation of persistent pollutants. These AOPs are based mainly on the in-situ production of hydroxyl radicals (OH) generated from an oxidizing agent (H2O2 or O2) in the presence of a catalyst. Among the most studied AOPs, the Fenton reaction stands out due to its operational simplicity and good levels of degradation for a wide range of emerging contaminants. However, it has some limitations such as the storage and handling of H2O2. Therefore, the use of the electro-Fenton (EF) process has been proposed in which H2O2 is generated in situ by the action of the oxygen reduction reaction (ORR). However, it is important to mention that the ORR is given by two routes, by two or four electrons, which results in the products of H2O2 and H2O, respectively. For this reason, current efforts seek to increase the selectivity of ORR catalysts toward the 2e route and thus improve the performance of the EF process. This work reviews catalysts for the Fenton reaction, ORR 2e catalysts, and presents a short review of some proposed catalysts with bifunctional activity for ORR 2e and Fenton processes. Finally, the most important factors for electro-Fenton dual catalysts to obtain high catalytic activity in both Fenton and ORR 2e processes are summarized.

1. Environmental Problems

One of the largest problems facing society today is water pollution, so special interest has been aroused in knowing the main pollutants that affect it and their subsequent treatment. Currently, some pollutants considered conventional (metals, fats and oils, nitrates, etc.) are controlled and monitored; however, more and more attention has been paid in recent years to those termed today as emerging [1,2,3]. Emerging contaminants can be defined as substances and/or chemical compounds that are not normally monitored but nevertheless remain continuous and remain present in water bodies despite being treated by conventional wastewater treatment plants [4,5,6,7].
Pharmaceuticals, along with their metabolites and degradation products, are considered emerging contaminants [8]. Their continuous discharge, together with their toxic and harmful levels, makes them highly dangerous for human and aquatic organisms [9,10]; therefore, it explains the importance of exercising controls and monitoring the concentrations of these pollutants both in the ecosystem and in effluents [11,12]. Currently, the global pharmaceutical market exceeds one trillion US dollars [13], which demonstrates the strong growth of this industry and its impact on the environment.
Among pharmaceutical products, antibiotics are of special interest because they are made to affect microorganisms, so they are prone to affect microbial communities in aquatic systems, causing a possible disappearance of them, which may have key functions in the ecosystem [14]. An antibiotic is defined as a chemotherapeutic agent capable of stopping the growth of microorganisms through cell destruction [15]. Antibiotic use has grown steadily globally, mostly in low- and middle-income countries [16]. It has been found that the presence of antibiotics in the environment favors the proliferation of microorganisms resistant to this type of drug [17].
Aquatic environments have been considered large deposits of antibiotics since this type of contaminant can arrive through two main ways, wastewater discharge and agricultural activity [18]. An example of this was a study conducted in 165 rivers in 72 countries, where researchers detected the presence of antibiotics in more than 60% of the 711 sampling sites [19]. In addition to these bodies of water, groundwater has also been susceptible to contamination with antibiotics mainly through the leaching of soils fertilized with manure from livestock, since it contains high concentrations of veterinary drugs [20].
Currently, wastewater treatment processes are composed of three phases in a general way:
  • Primary treatment by physicochemical operations;
  • Secondary treatment by biological processes;
  • Tertiary treatment by additional processes.
However, these treatment technologies only achieve moderate to significant elimination of some antibiotics [21]. For this reason, other types of treatments have been explored to achieve almost total elimination of the antibiotics present in water. Currently, in this study, advanced oxidation processes (AOPs) are explored [22,23,24], such as Fenton, electro-Fenton, ozonation, photocatalysis, and electrocoagulation, among others [25,26,27,28,29], which are based on the in situ production of hydroxyl radicals (OH) generated from an oxidizing agent (H2O2, O2) in the presence of a catalyst. OH has strong oxidizing power and thus, a high capacity to degrade organic compounds to simpler and less toxic biodegradable substances, and in some cases, even reaching complete mineralization [30].

2. Fenton Reaction

The first report on the Fenton reaction dates from 1876, where a mixture of H2O2 and Fe2+ salt (Fenton reagent) was used to achieve the oxidation of tartaric acid [31]; from there, various investigations have been carried out to apply this reaction in multiple applications. It is currently mainly used in environmental remediation [32,33,34,35,36].
The Fenton reaction can be considered an AOP, where it is based on a mixture of ferrous ion (Fe2+) (usually in a liquid state) with hydrogen peroxide (H2O2) [37], which, under the right reaction conditions, results in the formation of hydroxyl radicals [38] with high oxidative potential capable of degrading a wide range of contaminants (Figure 1) [39].
F e 2 + + H 2 O 2 F e 3 + + O H + O H
F e 3 + + H 2 O 2 F e 2 + + H + + H O 2
The Fenton reactions are based on the oxidation of Fe2+ to Fe3+, which simultaneously generates the hydroxyl radical (Equation (1)), followed by the reduction of Fe3+ to Fe2+ (Equation (2)), which allows the regeneration of the catalyst. However, the reaction in which OH is produced is much faster than that of catalyst regeneration [42], leading to the accumulation of Fe3+ and therefore decreasing catalytic efficiency.
Among its main advantages are its operational simplicity, zero toxicity, mild reaction conditions (ambient temperature and atmospheric pressures), and high degradation efficiency [43,44,45,46]. However, there are still some important limitations to overcome such as the slow reduction rate of Fe3+ to Fe2+ resulting in high levels of sludge generation with iron content, the deactivation of the catalyst, and a narrow pH working range [47,48,49,50,51].
The pH level is one of the most important parameters because it limits degradation efficiency. At pH values below 3, H2O2 goes on to form oxonium ions (H3O2+), which have less oxidative power than OH, while as the pH rises above 3, colloidal ferric species such as ferric hydroxide are formed from dissolved iron, which produces high amounts of sludge [52].
Recently, numerous investigations have been carried out (see Table 1) focused on the optimization of Fenton conditions. Variables such as catalyst pollutant and H2O2 concentration have been evaluated, as well as the pH of the solution, to determine under what conditions the highest percentage of pollutant molecule degradation is reached.
Typically, the pH is a crucial factor in the Fenton process, hardly affecting catalyst stability and pollutant degradation. From this, numerous investigations have been carried out with different pH values found at important degradation levels of pollutants such as antibiotics, colorants, and industrial wastewater. It is known that the pH optimal value for the Fenton reaction is closer to 3 due to the stability of H2O2, OH, and Fe3+. In this sense, Rao et al. [53] treated complex recalcitrant wastewater using FeSO4·7H2O as a catalyst (typical Fenton catalyst) in a concentration of 50 mg L−1 and a pH value of 3 obtaining a degradation of 72% of DQO. On the other hand, Zhao et al. [54] removed a polymer quaternary ammonium salt using the same catalyst but using a pH of 2 and a Fenton catalyst concentration slightly lower (40 mg L−1), achieving a percentage of elimination closer to 70%. Liu et al. [55] degraded a reference colorant for degradation tests, methylene blue, and analyzed the activation of the Fenton reaction under visible light irradiation using MoS2-Fe (150 mg L−1) as the Fenton catalyst at a pH value of 3, reaching a degradation of 65%. In turn, Qin et al. [56] achieved a 99.1% degradation level for a widely used antibiotic, amoxicillin, using 1 g L−1 of a core-shell structured MnFe2O4@C-NH2 at a pH of 3 (Figure 2).
Other antibiotics tested at pH 3 have been sulfamerazine (40 mg L−1), sulfathiazole (0.5 mg L−1), azithromycin (0.1 mg L−1), norfloxacin (50 mg L−1), and sulfamethoxazole (0.01 mg L−1), with removal percentages >65%, using as Fenton catalysts, CNTs-Fe3O4 (0.5 mg L−1), dissolved iron from zero-valent iron nanoparticles (9 mg L−1), MnFe3O4-HS (1000 mg L−1), Fe3O4/Schwertmannita/carbon (50 mg L−1) (Figure 3) and fresh powder from Fe0, respectively [57,58,59,60,61]. On the other hand, Hommem et al. [62] studied a range of pHs from 3.5 to 4.5, for the degradation of amoxicillin (0.45 mg L−1) using FeSO4·7H2O (0.095 mg L−1) as catalysts, obtaining 100% degradation in all cases.
Other authors tried to increase the operating pH to facilitate the use of the Fenton process in a real application. Cheng et al. achieved a tetracycline hydrochloride degradation of 91.3% at a pH value of 4 using magnetic pullulan hydrogel as a Fenton catalyst (1250 mg L−1) [63]. Similarly, Wang et al. [64] obtained a good tetracycline degradation (83%) at a pH value of 4.3 using iron loaded in graphitic carbon derived from microplastics. The highest pH values (5 to 5.5) tested also obtained good results: 100% degradation of carbamazepine and norfloxacin and 86.9% for ciprofloxacin using LaCu0.5Mn0.5O3, FeSO4·7H2O with hydroxylamine (10 μM of Fe2+), and δ-Fe3OOH/MWCNTs as Fenton catalysts, respectively [65,66,67]. However, neutralization of the final effluent is still required, so the search for catalysts and conditions that allow the development of this process at a neutral pH value is preferred. In this way, several researchers have studied the degradation of different pollutants such as amoxicillin, sulfadiazine, tetracycline, sulfadiazine, and rifampicin metronidazole, among others, working at pHs of 6 to 7.5, and obtaining degradation higher that 90% using goethite, Fe2(SO4)3·H2O, single-atom iron fixed in porous carbon, copper-modified MgFe-CO3 double-layered hydroxide, rGO@nFe/Pd, magnetic biocarbon, CoFe2O4-S, and LaCu0.8Mn0.2O3 [68,69,70,71,72,73,74,75] as Fenton catalysts. Guo et al. [76] also achieved the total degradation of ofloxacin at the widest pH using GO–Fe3O4 in a discharge plasma system. Wang et al. [77] also obtained the total degradation of ofloxacin using Fe3O4@S-doped ZnO at a pH of 5.2 to 9.0. For their part, Sun et al. [78] worked in a wide range of pH values from 3 to 9, degrading 99% of sodium sulfadiazine with FeCO3 [6000 mg L−1].
Table 1. Operating conditions for some Fenton processes.
Table 1. Operating conditions for some Fenton processes.
CatalystCatalyst ConcentrationH2O2 ConcentrationPollutantPollutant
Concentration
Solution
pH
Percentage of EliminationRef.
LaCu0.5Mn0.5O30.6 g L−10.7 g L−1Carbamazepine (CZP)15 mg L−15.5100%[65]
ZnO dopped Fe3O4@S0.25 g L−15 mL L−1Ofloxacin (OFX)10 mg L−15.2–9.0100%[77]
FeSO4·7H2O50 mg L−126.4 mMIndustrial wastewater (DQO)850 mg L−1 (DQO)372%[53]
FeSO4·7H2O40 mg L−123 mL L−1Hydroxyethylpolydialyldimethylammonium-acrylamide-acrylic-acrylate (PDM)200 mg L−1 269.3%[54]
Siderite (FeCO3)6 g L−1100 mMol L−1sodium sulfadiazine50 mg L−1999%[78]
Goethite-460 mg L−1Amoxicillin105 mg L−16.5–783%[68]
MnFe2O4@C-NH21 g L−13 mL L−1Amoxicillin30 mg L−13.099.1%[56]
Fe2(SO4)3·xH2O30 mg L−1-Tetracycline50 mg L−16.090%[69]
Fe2(SO4)·7H2O0.095 mg L−12.35 mg L−1Amoxicillin0.45 mg L−13.5–4.5100%[62]
rGO–Fe3O40.23 g L−1-Ofloxacin20 mg L−17.0–4.099.9%[76]
MoS2-Fe150 mg L−110 mMol L−1Methylene blue25 mg L−1365%[55]
CNTs–Fe3O40.5 g L−124.5 mMSulfamerazine40 mg L−1370%[57]
Dissolved iron from zero valence iron nanoparticles (nZVI)9 mg L−134 mg L−1Sulfathiazole0.5 mg L−13>96%[58]
Fe2(SO4)·7H2O with hydroxylamine10 μM de Fe2+1.0 mMNorfloxacin10 mg L−15100%[66]
CuS@Fe3O4/Pt--Tetracycline40 mg L−1-78%[79]
Single-atom iron fixed in porous carbon (Fe-ISA@CN)100 mg L−110 mMSulfadiazine2 mg L−16.596%[70]
LDH–CuMgFe–CO30.5 g L−14 mMol L−1Sulfathiazole0.15 mg L−17.5100%[71]
rGO@nFe/Pd200 mg L−1167 mMolRifampicin50 mg L−16.1494.6%[72]
MnFe2O4–HS1 g L−129.4 mMAzithromycin0.1 mg L−13.092.6%[59]
MagFePC (Hybrid)250 mg L−15 mL L−1Tetracycline-6.0100%[80]
Magnetic pullulan hydrogels1.25 g L−11.25 mL L−1Tetracycline hydrochloride20 mg L−14.091.3%[63]
Magnetic biocarbon0.2 g L−150 mMMetronidazole20 mg L−16.097.4%[73]
CoFe2O4–S500 mg L−15 mMol L−1Tetracycline0.1 M7.090%[74]
COF
(Photo-Fenton)
167 mg L−150 mL L−1Rhodamine B150 mg L−14.441.2%[81]
Fe3O4/Schwettmann/Carbon50 mg45 μLNorfloxacin50 mg L−13.0100%[60]
δ-Fe3OOH/MWCNTs235 mg L−120.6 mMolCiprofloxacin10 mg L−15.386.9%[67]
Fe2+5.0 mMol L−1 de Fe2+10 mmol L−1Antibiotic resistance gene (ARG)11.53 Log10(copies) g−1 dry sludge366.86%[82]
Fe–MPC0.02 g L−11.0 mMTetracycline40 mg L−14.383%[64]
Fresh powder from Fe0--Sulfamethoxazole10 μg L−13100%[61]
LaCu0.8Mn0.2O30.2 g L−113.8 × 10−3 mol L−1Paracetamol50 mg L−16.790%[75]
Apart from the commonly mentioned conditions, relevance has been found in some unusual factors, such as the slow and continuous addition of diluted H2O2, which improved the stochiometric efficacy of the reaction compared to the concentrated reagent being added only once [83]. Similarly, a high concentration of H2O2 causes an unfavorable effect on the reaction by bringing OH to H2O [84], which decreases the efficiency of the process.
Although high percentages of elimination have been achieved for some pollutants, it is important to note that those tests in which dissolved iron has been used in solution (conventional process) have an optimal pH value close to 3, which maintains the previously mentioned limitations.
In order to overcome its limitations, one of the variations in the conventional process is the use of chelating agents, which can allow the reaction to happen at a pH close to neutral; however, the possible harmful effects such as toxicity and contribution to total organic carbon remain to limit its use [85]. Therefore, the use of heterogeneous catalysts is proposed, where the reactive species are generated mainly on the surface of the catalyst [86], and not, as in the homogeneous process, where the reaction happens in the solution. This makes the recovery and reuse of the heterogeneous catalyst easier [87] and decreases the leaching of Fe3+ and, therefore, the generation of sludge. However, some heterogeneous catalysts, although they present minor leaching, behave as if they were homogeneous [88].
An ideal Fenton heterogeneous catalyst must have high catalytic activity and stability, and be easy to recover and reuse [89,90]. To supply this, different materials have been evaluated with iron in different oxidation states. Fe0 has been reported as a possible heterogeneous Fenton catalyst when maintained in acidic conditions and in the presence of dissolved O2 [91,92]. In the presence of oxygen, the in situ generation of H2O2 in zero-valent materials is possible, since the surface of the material can reduce O2 using two electrons, accompanied by the generation of Fe2+ [93]. Iron oxides, possessing Fe in different oxidation states and being the most abundant minerals in the earth’s crust, have been evaluated in the Fenton process. Among these, goethite [94], hematite [95], and magnetite [96] have been investigated. Among the mentioned iron oxides, Fe3O4 has aroused greater interest, due to its zero toxicity, effective catalytic activity (thanks to its Fe2+ species in its crystal structure), as well as its easy recovery and reuse due to its magnetic properties [97,98,99,100]. However, Fe3O4 presents a low catalytic activity, mainly due to the slow speed of the Fe3+/Fe2+ cycle. Therefore, several strategies have been proposed to overcome this limitation, such as coating, use of materials as supports, adjuvants, etc. [101].
It has been found that the addition of different carbon materials can enhance the Fenton reaction because the presence of carboxyl, carbonyl, and quinone groups present in the carbon material surface accelerates the reduction of Fe3+ to Fe2+ (Figure 4) [102,103,104,105,106]. On the other hand, carbon materials have been also used as support of the Fenton catalyst. Sun et al. [107] prepared ferric hydroxyquinoline supported on active carbon fiber, which enhanced the Fe3+/Fe2+ cycle through the free electrons present in the carbon material. Yao et al. [108] supported copper ferrite on reduced graphene oxide finding that the presence of two metal ions improved the redox cycle of both ions Cu and Fe together, with a synergistic effect between the nanocube structure of the copper ferrite and the mesoporosity of the support. Kuśmierek et al. [109] and Cruz et al. [110] used organic xerogel as support for Fe/N and CoFe2O4 and, in both investigations, a catalytic improvement was observed and ascribed to a higher activation of H2O2 due to the coexistence of two differents atoms (Fe and N or Co and Fe) in a same matrix.
This bimetallic synergism and the effect of the support were also observed by other authors. Zhang et al. [111] prepared Fe/Cu bimetallic-phase-supported montmorillonite using a two-step strategy of impregnation and calcination and found excellent efficiency in the removal of rhodamine B in comparison to single-metal-loaded montmorillonite. These results were ascribed to (i) the expansion of the montmorillonite pore size due to the insertion of Cu and Fe atoms making this clay more suitable as a Fenton catalyst, (ii) the electrons involved in the redox reaction of bimetallic active sites promote the rate of mutual oxidation reactions by enhancing the H2O2 conversion into oxidant radicals, and (iii) the leaching of the pure metal ions decreases due to the good stability of support. Li et al. [112] synthesized Fe–Mn/SAPO-18 zeolite catalysts by ion exchange, maintaining the zeolite structure. The author observed an enhanced degradation efficiency of methyl orange compared to the single-metal catalysts (Fe and Mn), which was also attributed to the synergistic effect of bimetal that improves OH radical generation and also promotes the formation of another active species, HO2. Similarly, Xu et al. [113] used sepiolite as support for Fe3O4 dispersion to be used as catalysts for the Fenton degradation of bisphenol A (BPA). The high surface area of the catalyst and the electron transfer and π-π interactions are responsible for the good BPA adsorption properties. Thus, the preconcentration of BFA by adsorption near the active site centers and the generation of hydroxyl radicals in the reaction of H2O2 and active sites make the Fenton degradation easier on the surface of Fe3O4-Sep. Ferroudj et al. [114] tried to analyze the catalyst’s structure–activity relationships and the effect of Fe3+ doping in the Fenton degradation of methyl orange (MO) and reactive back 5 (RB5) by the synthesis of Fe3+ doped and undoped maghemite/silica microporous and mesoporous microspheres (γ-Fe2O3/SiO2 MS). For that study, the authors concluded that porosity and Fe dispersion play an important role in the Fenton reaction. The creation of mesoporosity in the support enhances the reactants and products diffusion allowing a faster degradation of the tested dyes regarding its microporous counterpart, mainly for the RG5 molecule due to the biggest size. This degradation activity was improved after the doping of the silica framework with Fe3+. Thus, the best results were obtained for the Fe-doped γ-Fe2O3/SiO2 MS mesoporous catalyst. This material allows an almost complete degradation of MO in 25 min in contrast to the doped microporous catalyst, which achieved 92% of MO degradation in 1 h, and the undoped materials, which produces a much low degradation efficiency (63%) in 1 h. Similar results were obtained in the degradation of RB5.
The above-commented strategies to enhance the Fenton active-phase performances and the improvements achieved are collected and summarized in Table 2.
As evidenced in the literature, supporting Fenton heterogeneous catalysts is an effective method to increase the speed of the Fe3+/Fe2+ cycle, making it possible to enhance the efficiency of the process. In the same way, an interesting option is the use of metal-doped (e.g., N–Cu) or bimetallic active phases (e.g., Fe–Mn or Fe–Cu), which not only favor redox or Fenton reactions but also more complex reactions such as the oxygen reduction reaction (ORR) (e.g., nitrogen), which is very important for the electro-Fenton process, another variable of conventional Fenton that is addressed below.

3. Electro-Fenton Process

Heterogeneous Fenton has been proven to be a process with considerable potential for treating a wide range of pollutants. However, the storage, transport, and handling of H2O2, as well as the high costs associated with the excess consumption of the reagent are a limitation of the process [117]. To overcome these problems, the heterogeneous electro-Fenton (HEF) has been proposed, where the main contributions are the in situ production of H2O2 through the ORR at the cathode, the reduction of Fe3+ in the same electrode to supply Fe2+ again and the improved generation of OH by the applied electricity [118,119,120].
This process has been successfully used for the degradation of different recalcitrant pollutants such as drugs, dyes, petroleum derivatives, etc. [121,122,123,124]. The mechanism of the electro-Fenton process is given by several stages, summarized in [125]:
  • Electrogeneration of H2O2 (cathodic reduction of O2).
  • Production of OH (Fenton reaction).
  • Promotion of the formation of OH physisorbed on the surface of the electrode.
  • Regeneration of Fe2+ through the direct reduction of Fe3+ at the cathode.
In general, the reactions involved in the process can be summarized in Equations (3)–(6) [126,127,128]. However, it is necessary to pay special attention to the possible side reactions associated with the electrolysis of water that can decrease the performance of the catalytic activity of the process [129].
F e 3 + + e F e 2 + ( C a t h o d e )
O 2 + 2 H + + 2 e H 2 O 2 ( C a t h o d e )
2 H 2 O 4 H + + O 2 + 4 e ( A n o d e )
F e 2 + + H 2 O 2 + H + F e 3 + + H 2 O + O H
One of the most important benefits of this process to highlight is the expansion of the working pH range with the removal of contaminants at a pH close to neutral as well as its improvement in levels of biodegradability and toxicity [130,131]. However, as evidenced by Equations (4) and (6), ORR is the limiting step of the process, so it is important to pay attention to the cathode used [132,133,134].
Traditionally, carbon materials have been used as cathodes due to their good conductivity, stability, and low cost. Thus, these carbon materials are presented as excellent candidates as high-performance cathode electrodes for electro-Fenton reactions for several reasons that are described in the following. Sun et al. [135] synthesized a family of porous biochars by the chemical activation of cellulose powder with ZnCl2 and demonstrated that the carbonization temperature plays an important role in the catalytic performance in the EF oxidation of organic contaminants. While a high carbonization temperature favors the generation of water via the 4-electron pathway, a mild temperature of carbonization (400 °C) improved the generation of H2O2 via the 2-electron pathway. This different behavior is attributed to the surface chemistry of the carbon surface; mild carbonization temperatures provide an oxygen-functional-group-rich surface with a high C-O/C=O ratio, which enhances the H2O2 selectivity. In contrast, the use of higher carbonization temperatures (700 °C) guarantees a highly graphitic structure and created a carbon defect-rich porous structure highly selective to the 4-electron route. As a consequence, the biochar prepared at 550 °C presents a better catalytic performance as the cathode material for the EF degradation of several organic pollutants in presence of Fe (II), presenting a high H2O2 production rate (796.1 mg/g/h at −0.25 V vs. RHE at pH of 1) and achieving an almost complete removal (99%) of phenol after 1 h (at −0.25 V vs. RHE, pH = 3, [Fe (II)] = 0.5 mM, electrolyte: ((Na2SO4) = 0.1 M). Ergan et al. [136] also analyzed the effect of the thermal activation of activated carbon fibers in the H2O2 electrogeneration and EF efficiency. In this case, the effect of the carbonization time (from 1 h to 10 h) at 900 °C and also the carbonization atmosphere (CO2 or N2) was analyzed, observing that the best H2O2 production and dye (acid orange 7) degradation (92.9%) was obtained when the fibers were activated in an N2 atmosphere for 5 h due to a superior hydrophobicity, crystallite size, and larger electroactive surface area. Ramírez-Pereda et al. [137] compared two types of carbon materials: carbon fibers (CF) and vitreous carbon (VC) as electrodes for the electrogeneration of H2O2. The authors found that the generation of H2O2 is enhanced using CF (96% of H2O2 production efficiency) in comparison to VC (54% of H2O2 production efficiency), which was attributed to possible competitive reactions. Xia et al. [138,139] analyzed the performance of a polyacrylonitrile-based carbon fiber brush (PAN–CFB) as a cathode electrode for ORR reduction to H2O2 and studied the effect of the electrochemical modification of its surface. For the unmodified PAN–CFB, the highest production of H2O2 (current efficiency > 90%) was observed at a potential of −0.9 V in an acidic medium, whereas for electrochemically modified PAN–CFB, it was obtained at −0.4 V. However, in a neutral medium (pH = 7), an electrochemically modified PAN–CFB cathode inefficiently generated H2O2 (41.2% of current efficiency) in comparison with unmodified PAN–CBF, which still retains the high current efficiency (>90%). This was explained based on the surface chemistry modification after the electrochemical modification treatment, which converts the N-containing groups from pyridinic-N to 2-pyridone—these newly ORR active sites were influenced by solution pH, leading to different pathways in acidic or neutral media. Gao et al. [140] synthesized porous carbon felt modified with polypyrrole (Ppy) and anthraquinone 2-sulfonate, which had a higher specific surface area with more active sites for ORR; in addition, the modification with Ppy decreases the charge transfer resistance. In turn, Chen et al. [141] showed that the adsorption capacity of the carbon-based cathode plays an important role in the EF degradation of tetracycline. For that, they activated carbon felt with KOH at different temperatures, finding that the surface area and the oxygen-containing functional groups on the surface increased in the KOH activation, which significantly enhanced the drug adsorption capacity (being the sample carbonized at 900 °C, which presents the highest adsorption capacity) and also produced the best tetracycline degradation. This highest degree of pollutant degradation (tetracycline 80% removal) was explained based on the 3D structure and good adsorption capacity of activated carbon felt, which creates a space where the organic pollutant is adsorbed and preconcentrated near the active sites in charge of generating the H2O2, and, consequently, exposed to a high concentration of hydroxyl radicals in the system of homogeneous catalysis. Wang et al. [142] obtained a complete degradation of dimethyl phthalate using KOH-activated graphite felt at 900 °C. This high performance was attributed to the high surface area, the presence of oxygenated functional groups, and the more hydrophilic surface generated by the activation of the KOH.
Want et al. [143] pointed out the importance of electrode conductivity on its EF performance. For that, CNT and graphene-modified carbon felt were synthesized and the effect of this modification was analyzed in the degradation of azo-dye (RB5) by EF reaction. The modification of the carbon felt with CNT and graphene increases the degradation rate by 1.2 and 1.5 times than with the original carbon felt, respectively, giving degradation rates of 55.3% and 70.1%, respectively, which was attributed to the improved roughening of the smooth carbon felt fiber (specific surface area), conductivity, and corrosion resistance.
Since the ORR reaction is crucial for the H2O2 production, the improvement in the O2 supply to the active sites seems to be a key factor for the improvement in the EF efficiency, together with, as described above, the preconcentration of the pollutant near the H2O2-producing active sites. In this sense, Özcan et al. [144] compared carbon felt with carbon sponge as cathode materials for the degradation of basic blue 3 (BB3) by EF process finding a faster BB3 degradation using the carbon sponge-based electrode, which was associated with the major H2O2 production. Sopaj et al. [145] also evaluated the performance of carbon sponges in comparison with carbon felt and stainless steel in sulfamethazine degradation. The highest production of H2O2, and, consequently, the highest degradation, was obtained using the carbon sponge instead of carbon felt, the classical cathode for the electro-Fenton process, and stainless steel. This improved catalytic performance of carbon sponge is attributed to the high specific surface are and porosity, which results in inadequate conditions for the mass transport of species of interest (O2 and H2O2) throughout its 3D structure. Ganiyu et al. [146] also synthesized a carbon foam with high performance (total degradation and high mineralization) in the EF degradation of sulfanilamide, which was also attributed to the interconnected spherical cells’ porous structure (Figure 5), with high-area B.E.T. that improved the diffusion of substances and gaseous oxygen in the pores and active sites of the cathode, together with an excellent hydrophobicity and conductivity. Nonetheless, it exhibited poor performance in the electro-regeneration of Fe2+ from the reduction of Fe3+. This poor Fe2+ electro-regeneration of “gas diffuser materials” was also observed by Brillas et al. [147]. They demonstrated that H2O2 generation is important; however, it is not a uniquely important factor in the degradation of pollutants by EF. They analyzed the Fe(III)–EDDS-assisted EF degradation of butylated hydroxyanisole (BHA) using a carbon felt or an air-diffusion electrode as a cathode. They observed a much higher Fe(III) reduction efficiency in the novel Fe(III)–EDDS-assisted EF process in comparison with conventional EF. Moreover, the carbon felt resulted in a higher butylate hydroxyanisole degradation despite having an H2O2 generation lower than the air-diffusion cathode, which was attributed to the greater regeneration of Fe2+ observed using the carbon felt. To overcome this limitation, Chu et al. [148] used a dual cathode system: a gas diffusion electrode (GDE) to generate H2O2 by O2 reduction and a graphite electrode for the regeneration of Fe2+ from Fe3+ and analyzed its performance as a conventional single-cathode system in the degradation of 4-nitrophenol. They observed that the dual cathode system led to a more effective degradation of 4-nitrophenol, even with a lower initial Fe2+ concentration attributed to the rapid change between the Fe3+/Fe2+ couple in the dual cathode system. Since the main function of GDE is the electrogeneration of H2O2, the regeneration of Fe2+ by the cathodic reduction in the single-cathode system is very weak, whereas this regeneration was enhanced by the reduction at the graphite cathode in the dual-cathode system. Chu et al. [149] have also improved the reduction of Fe3+ of graphite by its modification with acidified carbon nanotubes, obtaining a degradation of 94.7% of p-nitrophenol. The acidification of the carbon nanotube enhanced the performance in the Fe3+ reduction in comparison with the unmodified graphite cathode.
Other advanced electrodes that improve the O2-supply are gas diffusion electrodes (GDE) evaluated by different researchers. Zarei et al. [150] synthesized a three-dimensional graphene (3DG)-based carbon paper GDE for the removal of nalidixic acid. The nalidixic acid removal was 90%, which was attributed to a higher surface area with more irregularities and roughness than graphene oxide, allowing rapid diffusion of O2 to achieve higher H2O2 production. On the other hand, Zhou et al. [151] used graphite felt, but in this case, as a floating cathode in which one side of a porous cathode is open to the air, and the other side is submerged in the aqueous solution instead of the traditionally submerged cathode. This novel strategy allows the creation of a three-phase boundary (O2 gas, electrolyte, and cathode) allowing the usage of O2 from both air and the electrolyte. This disposition is more energy-efficient and cost-effective in comparison with a GDE because avoids hydrophobic gas diffusion layer formation and the use of an air or O2 pump. An ibuprofen degradation of 78.3%, with only 25.4% obtained in the conventional position. Liu et al. [152] used a GDE of graphene doped with N-doped carbon nanotubes as the cathode in the dimethyl phthalate removal (100%), finding that the graphene–nanotubes interaction significantly improve the ORR activity.
The hydrophobicity of the cathode ©s also so important in the EF reaction. Chu et al. [153] managed to completely degrade cefepime with a sandwich-like superhydrophobic carbon cathode composed of graphite, carbon nanotubes, and PTFE observing that the hydrophobicity of the electrode significantly affects the ORR to H2O2 (Figure 6). Moreover, the prepared superhydrophobic carbon cathode resulted in an excellent reduction of O2 to H2O2, and, consequently, an excellent cefepime degradation, which is mainly dependent on the ∙OH production via Fenton reactions. Karatas et al. [154] also attributed the excellent performance of a carbon black electrode for the full degradation of atrazine to the superhydrophobic character of the electrode obtained because this high hydrophobicity provides high H2O2 yield by supporting the mass transfer of oxygen molecules.
Another strategy to improve EF performance is the use of metallic-doped electrode materials. Huang et al. [155] prepared a graphite felt-based cathode by first doping the graphite felt with Fe–Mn, then with active carbon, carbon black, and PTFE, and used it for the degradation of ciprofloxacin (Figure 7). This modified graphite felt-based cathode showed a higher degradation efficiency (95.4%) and wider operational pH range than their unmodified counterpart, which was principally associated with the larger surface area and volume pore as well as more active sites. Various researchers have also proposed doping carbon materials for improving EF efficiency. Ma et al. [156] doped multiwalled carbon nanotubes with Ag and Cu foam and noted that the high Ag dispersion improves the ORR and Fe2+ regeneration. Dung et al. [157] synthesized cobalt ferrite-coated carbon felt for the degradation of tartrazine (97%), where it is noted that the presence of Co3+/Co2+ and Fe3+/Fe2+ pairs improve the pollutant degradation; additionally, an additional Fenton catalyst was not needed in this test because the cathode fulfilled both functions. Paz et al. [158] used doped Vulcan XC72 with tungsten oxide nanoparticles and showed an improvement in the H2O2 production due to the surface being more acidic and hydrophilic. Fdez-Sanromán et al. [159] degraded 100% of pymetrozine with a cathode based on carbon felt and modified with carbon nanofibers doped with iron; the improvement was explained by the porous structure that was enhanced with the incorporated nanofibers, and it was also explained that the catalytic activity of a typical Fenton catalyst is more effective in the presence of iron.
In general, it could be summarized that the important parameters to take into account in the cathode preparation are the mesoporous structure that translates into the high surface area, the hydrophobicity that plays an important role in terms of the diffusivity of O2 inside and outside of the electrode, as well as the regeneration of Fe2+ that happens in the electrode, and finally, the composition of the cathode, especially the presence of efficient nitrogenous groups toward ORR [160,161]. However, more in-depth research should be conducted, despite the extensive literature that supports the above. In this regard, Cordeiro stressed that graphite within a series of carbonaceous materials presented the smallest B.E.T. area and zero existence of oxygenated groups on the surface, and still showed the best efficiency of H2O2, which was associated with its laminar structure and low hydrophobicity leading to the faster release of H2O2 compared to more porous materials [162].
Another important parameter to consider in the cathode environment is the balance between the three reactants (oxygen, proton, and electron), which must have a molar rate of 2:1:2, respectively [163]. For this, it is important to start studying how to supply air/O2, which is based on two types of aeration specifically.
  • Direct aeration in solution.
  • Aeration through a gas diffusing electrode (GDE).
Direct aeration in solution through bubbling is the traditional method; on the other hand, GDEs have the advantage that thanks to their porous structure and coexistence of the limits of three phases, they have solved the problems of low solubility and mass transfer of O2 [164], which can accelerate ORR kinetics and gas-use efficiency, summarizing even better energy use. In general, GDEs can improve ORR mainly due to two main mechanisms [165].
  • The increase in the mass transfer of O2 improves the contact of the gas with the catalyst and, therefore, the active sites for the ORR.
  • The increase in the mass transfer is not only in reference to O2 but also between the catalyst and electrolyte, which promotes the release of H2O2 and active sites, improving the formation of the oxidizing species.
GDEs can be defined as porous electrodes in which the solid part is in contact with the gas phase, and then with the solution to be treated [166], which generates a three-phase interface in which the reactions of interest will happen. GDEs are mainly composed of a catalyst layer (in contact with the solution) and a diffuser (in contact with air or oxygen) (Figure 8) [167]. At this point, it is important to clarify that a very thick layer of catalyst can decrease the output rate of H2O2 to the solution, which can cause it to degrade on the surface of the material. An alternative to adjusting the thickness of the layer or avoiding its limitation is to adjust the hydrophobicity of the catalyst [168].
As it has been pointed out, good results have been obtained in the EF degradation of a huge variety of emerging pollutants. However, since H2O2 production by the O2 reduction at the cathode seems to be the limiting step, a deeper investigation into ORR is necessary to better understand how cathode variations benefit or affect the efficiency of the EF process.

4. Promising New Fenton Catalysts: Spinel Ferrites and Perovskites

Fenton catalysts with higher reaction rates are currently being sought to increase their overall efficiencies. Among these, spinel ferrites have attracted increasing interest in recent years due to their low cost, excellent catalytic activity, and magnetic properties that allow a facile separation and reutilization. In this section, the synthesis routes, advantages, and applications of spinel ferrites will be described.
There are different methods of synthesis within which the following stand out:
  • Chemical coprecipitation
Chemical coprecipitation is one of the most promising methods thanks to its ease of realization, as well as the possibility of mass production. It is mainly based on the co- precipitation of materials by a pH change. In general, this method consists of a mixture of precursors with the species of M2+ and M3+ (metal), to later precipitate the particles by the addition of a base (KOH, NaOH, or NH4OH). It can be summarized in Equation (7) [169,170,171,172].
M 2 + + 2 F e 3 + + 8 O H M F e 2 O 4 + 4 H 2 O

4.1. Sol–Gel Method

This method is one of the most used for the synthesis of nanoparticles due to its relatively low cost, operational simplicity, and high homogeneity of the material obtained. It is based on a series of reactions, summarized in the hydrolysis of a metal alkoxide and the subsequent polycondensation of the hydroxyl groups formed, which produces a three-dimensional matrix [173,174,175], which must subsequently be subjected to thermal processes to improve the crystallinity of the nanoparticles.

4.2. Hydrothermal/Solvothermal

This process can be defined as a series of chemical reactions that occur in a closed system with one or more precursors in the presence of the solvent (water for the hydrothermal case), at a temperature above the boiling point of this [176]. This process has several advantages such as its operational simplicity, versatility, and low cost, among others [177,178]. An important advantage to highlight is the well-controlled diffusivity within the system [179], which allows good control of the structure and morphology of the synthesized particles.
From these synthesis methods, it has been possible to obtain different new Fenton-type catalysts at the nanometer scale, with different coatings and supports [180,181,182] in order to improve the efficiency of the hetero-Fenton process. It has been established that in the hetero-Fenton process, having a catalyst with a higher amount of transition-metal ions and a higher specific area leads to better Fenton activity [183,184]. Fe3O4 is currently being extensively investigated as a Fenton catalyst [185,186,187] due to its relatively high activity and easy magnetic separation. An essential advantage is the presence of Fe2+ and Fe3+ in a single material, due to its cubic structure where half of the Fe3+ ions occupied all of the tetrahedral sites and the Fe2+ ions are founded in the octahedral, allowing the presence of the two important species in Fenton processes. It has been reported in the literature that transition-metal ions occupying tetrahedral sites are catalytically inert, while those located at octahedral sites tend to determine catalytic activity, which is due to effect that the metal cations in these positions are found exclusively on the surface of the solid and thus take part in reactions to generate OH radicals [188]. It has been suggested that the activation of H2O2 in the presence of magnetite takes place on the surface of the solid, i.e., the Fe3+-OH [189].
According to crystalline field theory, common metal cations have the following order of preference to occupy octahedral sites, Cr3+ > Ni3+ > Cu3+ > Al3+ > Mg2+ > Fe2+ > Co2+ > Fe3+ > Mn2+ > Zn2+ [190]. The most studied are Fe, Mn, Cu, Ni, Zn, and Co [191,192].
Spinel ferrites have a general formula of A2+B3+2O−24 and a cubic lattice structure, where positions A and B are occupied by divalent and trivalent metal cations, according to the distribution of cations in the lattice, and the spinels can be normal, random, or inverse. In a normal spinel, tetrahedral sites are related to position A and octahedral sites to B. In a reverse spinel, tetrahedral sites are occupied by half of the B cations and the octahedral sites by all of the A cations. This distribution of cations is responsible for determining the magnetic properties of ferrites [193]. There are different types of spinel-structured materials; however, in this work, we will focus specifically on three of them: copper and cobalt ferrites and iron cobaltite.
Copper ferrite (CuFe2O4) is a well-known material with a reverse-spinel structure, which has a stable structure that reduces metals leaching, together with unique magnetic, electrical, physical, and chemical properties [194,195,196,197]. These properties make them a promising Fenton-type catalyst. Cu+ ions have been reported to have the ability to generate OH radicals by a mechanism similar to that of Fe2+ according to Equations (8)–(10) [198].
C u + + H 2 O 2 C u 2 + + O H + O H
C u 2 + + H 2 O 2 C u + + H + + H O 2
C u 2 + + H O 2 C u + + H + + O 2
Feng et al. [199] synthesized CuFe2O4 nanoparticles to be used as a Fenton catalyst in the degradation of sulfanilamide. They suggest that CuO is more reactive and effective than Fe3+ for the activation of H2O2 and, which is more important, can work in a higher pH range than conventional iron oxides. Moreover, Zhang et al. [200] observed that the leaching of Cu2+ is 30 times lower in CuFe2O4 than with CuO. The promising performance of the copper spinel ferrite as a Fenton catalyst was pointed out by several authors. Suraj et al. [201] synthesized CuFe2O4 by the chemical coprecipitation method and used it as a heterogeneous Fenton catalyst for the treatment of pulp and paper wastewater, obtaining a 78% elimination of the chemical oxygen demand. Ding et al. [202] demonstrated that the morphology of the spinel is also very important. They synthesized hollow CuFe2O4 spheres with oxygen vacancies (Figure 9), which demonstrated greater degradation of ciprofloxacin than normal CuFe2O4 particles. This better performance of hollow spheres was attributed, among other factors, to the synergistic oxygen vacancies and confinement effects on the catalyst surface. The oxygen vacancies produce highly active electron-rich Cu+ species, which enhanced the H2O2 activation and, thus, the hydroxyl radical generation. In turn, the hollow structure is responsible for concentrating the organic pollutants near the OH-generator active sites, improving the organic pollutant molecules/OH radicals contact, and accelerating the degradation. According to Deyou et al. [203], the particle size and surface area are more important factors than a crystalline structure for improving the catalytic efficiency of CuFe2O4. López-Ramón et al. [204] evaluated the effect of calcination temperature on the catalytic activity of CuFe2O4 synthesized by the sol–gel method, finding that the calcination temperature has two opposite effects: the activity decreases with increasing temperature due to the increase in crystalline size and cubic-to-tetragonal transformation of ferrite and appearance of hematite; however, the metal leaching decreases with increasing calcination temperatures.
Cobalt ferrite (CoFe2O4), like other ferrites, has a high catalytic activity, stable crystal structure, low solubility, magnetic properties, as well as the ease of controlling the leaching of cobalt due to the strong interactions between metals, and the strong redox activity of Co with catalytic properties similar to those of noble metals (Pt, Ir, and Au) [205,206,207,208], which make it a promising material as a Fenton-like catalyst.
Feng et al. [209] synthesized monodispersed CoFe2O4 nanoparticles by a solvothermal method to evaluate them as a Fenton catalyst in the degradation of methylene blue, reaching a concentration decrease of 96.8%. The authors highlighted that the existence of Co2+ ions could favor the decomposition of H2O2 to OH and subsequently give way to different reactions, as evidenced in Equations (11)–(14).
C o 2 + + H 2 O 2 C o 3 + + O H + O H
O H + H 2 O 2 O O H + H 2 O 2
F e 3 + + O O H F e 2 + + H + + O 2
C o 3 + + O O H C o 2 + + H + + O 2
Sing and Singhal [210] demonstrated that the transition-metal doping of cobalt ferrites is a promising method for tuning the physical characteristics of catalysts and thus, enhancing their catalytic and magnetic properties. For that, the authors synthesized a series of Ru-doped cobalt ferrite nanoparticles by the sol–gel method for the photo-Fenton degradation of red Remazol textile dye, achieving a degradation of approximately 90% within 120 min. The mechanism proposed is based on the photocatalytic and Fenton character of the Ru-modified ferrite. An electron-hole pair is created by the irradiation of cobalt ferrite nanoparticles with visible light. The photogenerated electrons are responsible for the OH generation from H2O2 and also the reduction of the Fe3+ cation on cobalt ferrite to Fe2+, which further generates OH radicals in the reaction with hydrogen peroxide. Vinosha et al. [211] also analyzed the photo-Fenton performance of CoFe2O4 nanoparticles obtained by means of chemical coprecipitation, achieving almost total degradation of methylene blue (~99.3% in 75 min) under visible light irradiation. As an outstanding result, they proposed that the pH used in the synthesis was not an impact parameter that affected the morphology of the catalyst; however, it significantly affects the particle size (a more alkaline (pH > 9) medium, larger crystallite size). It has been proposed that the reactions that lead to the formation of CoFe2O4 by the chemical coprecipitation method in an aqueous medium, are those presented in Equations (15) and (16) [212].
2 F e 3 + + C o 2 + F e 2 C o ( O H ) 8
F e 2 C o ( O H ) 8 C o F e 2 O 4 + 4 H 2 O
In turn, Iron cobaltite (FeCo2O4) has been also studied in environmental remediation and energy storage, thanks to its electrical properties and electrochemical performance [213]. In the energy storage field, Mohamed et al. demonstrated that iron cobaltite nanorods show a better capacity and lower overpotential as the cathode of lithium−O2 batteries than other metal cobaltites (Mn, Ni, and Zn) [214] because the FeCo2O4 surface has the highest number of electropositive Co3+ active sites that improve the oxygen adsorption and Fe2+ in the tetrahedral site that favors the release of electrons to reduce oxygen. Yadav et al. [215] demonstrated that iron cobaltites are also efficient for supercapacitive and photocatalytic applications due to the valence states of the Fe3+/Fe2+ and Co3+/Co2+ species, which are considered active catalytic sites. These nanoflake-like iron cobaltites present a capacitance as high as 1230 F g−1 (5 mV s−1) with a good rate capability and superior cycling stability and also show a good photocatalytic performance achieving up to 94.19% degradation of crystal violet dye under sunlight illumination. However, despite this, very little work has been carried out with reference to their evaluation as a Fenton-like catalyst [216,217]. Zhang et al. [216] synthesized nitrogen-containing carbon/FeCo2O4 composites and analyzed their performances as Fenton catalysts for the degradation of methylene blue obtaining almost 100% removal in 10 min without pH adjustment, which was attributed to the uniform distribution of bimetals and nitrogen doping, which ensured the exposure of sites with high catalytic activity. Zhao et al. [217] analyzed the behavior of FeCo2O4/g–C3N4 as a photo-Fenton catalyst in the degradation of rhodamine B (RhB), obtaining 98% degradation in 45 min, which was attributed to a synergetic interaction between photocatalytic and Fenton-like reactions and the effective separation of the photogenerated charges. Therefore, it is proposed that the use of iron cobaltite as a pristine catalyst (without doping or support) in the Fenton reaction may be promising for future applications.
In general, spinel ferrites turn out to be attractive materials for catalytic activities in Fenton processes, mainly because they manage to improve the generation reaction of Fe2+ and achieve a synergistic effect between metal ions with valences similar to those of iron involved in the process (M2+, M3+, M: Metal).
Perovskites are other types of materials with promising catalytic activity that can be synthesized by the abovementioned methods (Chemical coprecipitation, sol–gel, and hydro/solvothermal) [218,219,220,221]. Perovskites can be defined as a type of mixed oxide with different formulations, binary (ABO), ternary (AA’BO or ABB’O), and quaternary (AA’BB’O), where A and B are cations sites occupied by alkali metals, alkaline-earth metals or rare-earth metals and transition metals, respectively [222].
Some perovskites have been studied in different Fenton-type reactions, for example, Carrasco-Díaz et al. [75] removed paracetamol by Fenton reaction using LaCu1−xMxO3 (M = Mn, Ti) as the catalyst and determined that the most active catalyst was the one that contained the highest amount of Cu2+ at the surface. Moreover, they found that the titanium and manganese species seem not to be responsible for the improvement of activity with respect to the sample LaCuO3. Li et al. [223] synthetized a Ca1−xFeO3−δ perovskite and determined that the A-site cation can distort the FeO6 octahedra in the perovskite and regulate the oxygen vacancies (OV) concentration; in this way, an A-site cation deficient of Ca0.9FeO3-δ results in an improved H2O2 activation for the degradation of tetracycline by a Fenton-like process. Similarly, Xie et al. [224] found that the copper incorporation in LaCoO3 perovskite improved the electro-Fenton activity due to the enhancement of redox activity and oxygen vacancies, but in this case, by the substitution of B-site elements, which synergistically promoted the activation of hydrogen peroxide to a hydroxyl radical (OH). On the other hand, Rusevova et al. [225] degraded phenol via heterogeneous Fenton-like reactions using iron-containing LaFeO3, and BiFeO3 perovskites, and made a comparison with data reported in the literature using, as a catalyst, nano-sized Fe(II, III) oxide particles, determining that the perovskites synthetized in this work had a higher catalytic activity. Zhao et al. [226] determined that BiFeO3 supported in carbon aerogel (BFO/CA) with a three-dimensional (3D) structure improves the catalytic activity and stability of BiFeO3, resulting in an interesting strategy for the development of advanced catalysts for its possible application in Fenton processes.
In summary, perovskites are materials similar to spinel ferrites, and the contribution of two different metal species (formation of OVs) can be of interest to the Fenton process.

5. Oxygen Reduction Reaction through Two Electrons (ORR 2e)

The oxygen reduction reaction (ORR) has attracted attention in recent years due to its applicability in various fields of industry, both in energy storage and water treatment processes [227,228,229]. However, this review is focused on ORR as a route for the electrogeneration of H2O2, a product that controls the efficiency of the electro-Fenton oxidation reaction.
In general, ORR involves two types of reactions [230]:
  • Oxygen molecules that are converted directly into water (4e);
  • Oxygen molecules that are initially reduced to peroxide and later to water (2e).
Analyzing the mechanism, it has been established that ORR 2e takes place in two steps involving two single proton-coupled electron transfers; in the first, there is the transfer of electrons and generation of the HOO intermediate linked to the active site (*), while in the second, H2O2 is produced (Equations (17 and (18)).
O 2 + * + H + + e O O H *
O O H * + H + + e H 2 O 2 + *
However, H2O2 types formed at the cathode can react in three ways [231]:
  • Being electro-reduced to H2O at the cathode (usually at porous cathodes);
  • Being oxidized to O2 at the anode via HO2· as an intermediate;
  • Disproportionate to O2 and H2O in a non-electrochemical reaction.
Therefore, the objective in the design of catalysts for the electroproduction of H2O2 by ORR is to synthesize materials capable of binding the OOH* intermediate with the appropriate strength; this is neither very strong nor weak to avoid the 4e pathway predominating [232].
The most studied way of improving the performance of the ORR 2e is the synthesis and optimization of the cathode materials, since, together with the generation of H2O2, the current efficiency and, therefore, the energy consumption needs can also be taken into account [233]. In general, an optimal catalyst for ORR 2e must have good electrical conductivity, stability, and resistance to acid corrosion [234], properties associated with carbon-based materials [235]. Carbon materials have proven to have excellent properties for use as a cathode due to their high electrical conductivity, zero toxicity, low cost, high stability, and high catalytic activity [236,237,238]. However, although carbonaceous materials exhibit high selectivity toward H2O2 production, they have a slow kinetic toward this reaction, resulting in poor performance.
Different strategies to improve the electrochemical production of H2O2 have been recently studied and one of them consists of the deposition of other catalysts or metal oxides (MnO2, V2O5, and CeO2) on the surface of carbonaceous materials [155]. On the other hand, doping with heteroatoms such as oxygen (O), sulfur (S), nitrogen (N), and fluorine (F) is another strategy developed to increase the electrocatalytic performance of carbon materials. The carbon lattice doping with heteroatoms introduces defects to the carbon structure, modifying their atomic and electronic structure and, thus, resulting in new promising materials for several energetic applications such as energy production and storage. However, the improvement mechanism of the introduction of these heteroatoms is still under debate by researchers worldwide. Some authors [239,240,241,242,243] found that the introduction of some oxygen functional groups can induce greater electrical conductivity and electrocatalytic activity thanks to the formation of a hydrophilic surface. Recently, different carbonaceous materials have been evaluated as ORR 2e catalysts, where the importance of the amount and type of oxygenated functional groups present on the surface of these has been evidenced, finding that a greater amount of carboxyl group translates into greater efficiency in H2O2generation because it facilitates the displacement of the electron density of the active site, allowing a better interaction in the adsorption of O2 [162,244]. In the same way, it has also been established that the higher the ratio of ID/IG in carbonaceous materials, the greater the production of H2O2 [245].
Zahoor et al. [246] indicated that the sp2 electron configuration of graphite, graphene, and nanotubes are rich in π electrons which can be transferred without resistance in ORR; however, this electron flow is not enough to promote ORR. Thus, the heteroatom introduction (e.g., N) within its network induces π electron activation by bonding with the solitary electron pairs of N and providing negative charges. The carbon atoms adjacent to nitrogen have a high positive charge density, which increases the adsorption of oxygen and reactive intermediates, which results in enhanced ORR activity. Okamoto et al. [247] found by density-functional calculations that the adsorption of O2 is more energetically favorable with the increase in the number of N around a C=C bond. Lagarreta-Mendoza et al. [248] proposed quantum phenomena for the explication of the ORR mechanism on N-doped carbon-based electrocatalysts, which consists of the hybridization change in carbon atoms in the graphene lattice, from sp2 to sp3. Nitrogen doping could induce a quantum tunnel phenomenon called nitrogen inversion, which consists of a process in which the lone pair of an sp3 nitrogen atom migrates from one face of the atom, travels through the nucleus (quantum tunneling), and reappears on the other side. This inversion of nitrogen (due to network defects) allows the N atom to act as a “switch” that activates or stops the flow of electrons, generating active sites for ORR The heteroatom–carbon interaction induces the activation of the electron π when joining with the lone pairs of N so that later, the O2 molecules are reduced into the carbons positively charged by the neighboring N. Additionally, it has been observed that the insertion of a single atom of N does not, apparently, have enough influence to alter the carbon, so a minimum of three atoms are necessary to cause small differences in the length of the C and N bonds [248]. As reported in the literature, N-pyrrolic is more likely to promote ORR 4e by increasing electron donation capacity, while N-pyridinic has more tendency for ORR 2e [249]. Wang et al. [250] investigated the ORR 2e activity of multiwalled carbon nanotubes (MWCNT) with different degrees of oxidation by oxidation with concentrated sulfate and potassium permanganate at 20 to 60 °C for 1 h. They observed that the presence of oxygen groups created after the oxidation treatment enhances the H2O2 generation, with the MWCNT always being oxidized and more active than the nonoxidized counterpart. Among the oxidized MWCNT, the sample prepared at middle conditions (40 °C and 1 h) present the best H2O2 selectivity which was explained by the composition of its outer and inner structure. After the oxidation treatment, the outer surface is damaged by the creation of a large number of defects together with the introduction of oxygen functional groups such as COOH and C-O-C that act as excellent active sites for the ORR 2e, while the inner structure is maintained, ensuring its good electrical conductivity.
Similarly, it is reported that carbon doping with heteroatoms and transition metals can energetically optimize the charge and rotation density of carbon, helping to promote the activation and reduction of O2 by weakening the O–O bond in the ORR process [251]. Some studies have found that N can change the mode of adsorption of O2, while sulfur (S) affects the binding capacity of the OOH intermediate by varying the spin density [252,253]. Ting Soo et al. [254] reviewed the electrochemical performance of modified graphene as an ORR electrocatalyst, and the spin density, charge, and energy gap were evaluated, finding that the microstructural properties of heteroatom-doped graphene and the high spin density or atomic charge density on adjacent carbon atoms are the key factors that contribute to the enhanced electrocatalytic activity. However, an over-dose of heteroatom doping has negative effects due to the formation of high-sized clusters that make electron transfer difficult. Zhang et al. [255] explained that any chemical species that is substituted for graphene can lead to a high asymmetry spin density and atomic charge density, which promotes ORR. However, by applying DFT calculations they demonstrated that ORR on N-doped graphene is a four-electron pathway. Conversely, Pei Su et al. [256] prepared N-doped graphene showing high H2O2 yield, selectivity, and low EEC. They also observed that the C=C and N-pyridinic bonds in N-graphene have the ability to enhance the onset potential, while the N-graphitic groups are responsible for the current disk obtained. Moreover, in EF reaction, they demonstrated that N-graphitic are responsible for the H2O2 generation, and N-pyridinic groups catalyze the OH radical productions from H2O2 (Figure 10). To further investigate the effect of the N-group amount and nature, Pei Su et al. [257] controlled the N-surface chemistry of N-doped graphene (N–GE) via pyrolysis temperatures (200, 300, 400, 550, 750, and 950 °C) in the catalytic activity, observing that the value of electrons transferred in the ORR was in the range of 2 to 2.5 for temperatures of 200 to 400 °C, and from 2.5 to 2.7 for further temperature increases. Thus, the sample pyrolyzed at 400 °C presented the highest active N content and H2O2 selectivity and achieved the highest phenol degradation (93.58%). They proposed that the conversion of graphite N and pyridinic-N in N–GE plays an important role in the oxygen reduction reaction (ORR) and OH conversion, while the conversion of pyridinic-N-oxide to pyridinic-N is critical for catalyst stability and sustainability.
Qin et al. [258] and Xie et al. [259] evaluated N and O co-doped carbon material, highlighting the synergistic effect of the different species composed of the two atoms in question on the generation of H2O2. The pyridine-N groups combined with ether (C-O-C)/carboxyl (COOH) were highlighted, which presented lower adsorption energy of the intermediate *OOH, facilitating protonation toward H2O2. Some studies [260] have also pointed out that the -COOH and carbonyl (-C=O) groups are active for ORR 2e, and mainly the -COOH group.
Iglesias et al. [261] synthesized N-doped graphitized carbon nanohorns (N–CNH) via coating and controlled annealing of polydopamine (PDA) (Figure 11) and analyzed its use as an electrode material for OEE 2e. The excellent catalytic performance of N–CNH to perform ORR 2e is explained based on the high surface area and accessible porosity of the CNH scaffold, a good distribution of different N-contained surface groups, and improved conductivity and easy electron transfer. The authors also highlight that the microporosity and the specific pyridinic/pyrrolic ratio are crucial factors affecting the ORR route. The microporosity decreases the residence time of H2O2 on the material avoiding further reduction to H2O2. The relative types of N atoms also affect the electrocatalytic performance. However, the influence is dependent on the pH. At an acidic pH, protonation of the N-pyridinic occurs which compromises the ability to weaken the O–O bond producing H2O2 as the main product while at a neutral or alkaline pH, a reduction in selectivity toward H2O2 is observed because the formation of H2O by the O–O breaking becomes progressively more available.
Chen et al. [262] also demonstrated that porosity is so important in controlling the selectivity of the process. Chen et al. determined that activated carbon with abundant micropores and a high surface area is better for high ORR activity via 4e, while carbon and graphite black with a smaller specific area tend to catalyze ORR via 2e. Granules with more micropores facilitated the ORR process and caused a further reduction of H2O2, leading to less H2O2 generation. Similarly, Wenhui He et al. [263] analyzed the effect of the 3D hierarchical porosity on N-doped carbon on the catalytic performance. For that, they prepared N-doped carbon with hierarchical micro-, meso-, and macroporosity by the use of 3D macroporous templates and endogenous pore-generating agents with the assistance of KOH activation. It was concluded that the catalytic efficiency of nitrogen-doped species in ORR highly depends on their degree of exposure to the carbon matrix and that the hierarchical structure can strongly influence the rate of diffusion of O2 and rapid release of H2O2.
In general, it can be concluded that there are some important requirements to obtain carbonaceous materials with high catalytic activity for ORR 2e such as high porosity (subject to further investigation), large surface area, presence of oxygenated groups, bonds C=C, N-pyridinic, and a greater amount of C atoms sp3 hybridization.
Zhang et al. [264] synthesized interface atomic domains (IAD) of C/N/O atoms (pyridine-N, C=O), from biomass tar by pyrolysis, oxidation, and doping. With this assured an ordered structure (by pyrolysis), with functional groups C=O (oxidation), which results as a host for the generation of N-pyridinic (doping). As a result, it was found that the material obtained had high selectivity for ORR 2e (N-pyridinic) and an excellent mass transfer of O2 (C=O and ordered structure).
Another strategy is the doping of porous carbon materials with transition metals which can accelerate the redox reaction by the action of metal ions, thereby increasing the rate of electron transfer. Similarly, nonprecious metals have proven to have an excellent capacity for ORR [265]. It has been shown that compared to bare carbon, metal/carbon compounds have a noticeable improvement toward ORR [266]. Metal/carbon catalysts can adsorb O2 in two ways, which translates into a greater or lesser efficiency toward reduction via 2 or 4 electrons [267,268,269,270].
  • Lateral adsorption leads to a longer bond length and therefore weakens the O=O bond, giving as a product, H2O (Figure 12).
  • Adsorption at the end, where oxygen is adsorbed in the form of OOH, which can produce H2O2 and H2O (Figure 12).
Thus, to obtain metal-doped materials selective to the electrogeneration of H2O2, the end-on adsorption of O2 molecules on the catalyst surfaces need to be promoted. One strategy is the isolation of the active atomic metal sites [271]. For that, three pathways have been developed: (i) alloying an active metal with an inert metal that can induce geometric isolation of the active metal where the O2 is end-on adsorbed; (ii) the partial poisoning of the active metal surface with an inert phase or molecules, e.g., the coating of Pt nanoparticles with amorphous carbon by CVD (Figure 13); and (iii) the synthesis of atomically dispersed catalysts.
Some nonprecious metals, used as a doping material in carbonaceous matrices, despite increasing costs, are offset by high dispersion in the matrix, which favors the production of H2O2 [273]. In general, although iron may be an ideal metal to regulate the binding energy of the intermediate *OOH with the carbon matrix, it has been shown that it has a strong tendency toward the route of 4e [274,275]. Co is one of the metals studied for ORR, demonstrating greater selectivity toward two electrons compared to Fe [276]. On the other hand, Chen et al. [277] investigated the incorporation of Cu atoms in mesoporous N/O-codoped graphitic carbon catalysts with a tailored mesoporosity concluding that the H2O2 production depends on the number of C atoms bound to the O atom, which increases by doping with a tiny amount of N. However, the inclusion of copper significantly decreased the selectivity toward H2O2. Cu, having a strong binding with oxygen, quickly generates blockages of the active sites due to the O and OH species generated in the reaction [278]. In other studies, Zhang et al. [279] synthesized Au@Cu2−xS-CNTs core-shell deficient in Cu and determined that the catalysts exhibited higher ORR activity and selectivity toward H2O2, associated with the Cu defects, which (from theoretical calculations) decreases the energy of reaction of OOH*. Wang et al. [280] probed that the nanostructure of the Cu active site is the key to the success of the catalyst; if the active site and host effect are optimally coordinated, it will certainly show considerable potential in the catalytic performance. For that, the authors designed electron-deficient and electron-rich Cu active sites by the regulation of nitrogenous ligand solvents and analyzed the effect in the capture and activation of the C≡C bond of acetylene. The results showed that the Cu sites, stabilized by N-pyrrolic, generate electron-deficient states that improve the activation capacity of acetylene and, thus, the catalytic activity (hydrochlorination). In turn, Hyo-Noh et al. [281] determined that Cu can strongly increase the chemical bonding with O (unfavorable for ORR 2e) and thus bare Cu particles cannot catalyze the ORR at all, but this energy can be easily controlled by the coating with N-doped carbon shells (regulated through CO2 treatment), resulting in ORR activity comparable to the Pt (111) surface.
Therefore, recently, the investigate into other types of metals has begun. Lenarda et al. [282] investigated different metal (Fe, Ni, Mn, and Co) nanoparticles embedded with an N-doped graphitic carbon and determined that the nature of the metal dictates the type of final surface N-group by obtaining different N-pyridinic/N-pyrrolic ratios, which is essential for H2O2 selectivity. The optimal ratio was obtained when Co was used.
Thus, Co is presented as a promising metal to increase the ORR activity without losing H2O2 selectivity. In a recent study, Gao et al. [283] prepared single-atom transition metals anchored in nitrogen-doped graphene citing that Co could be an active site for the ORR 2e pathway since the bond with the OOH* intermediate has optimized adsorption energy, neither too strong nor too weak, in addition to presenting a higher OOH* to O* reduction barrier than that achieved with iron and magnesium atoms. However, Cao et al. [284] evaluated Co nanoparticles supported on N-doped holey graphene aerogel in ORR, determining that the reaction pathway was given by a 4-electron process. Collman et al. [285] studied the geometric effect of cobalt on two different porphyrin structures (monomeric and dicobalt), where the monomeric consists of a Co atom surrounded by N, with which they were able to determine that the selectivity toward H2O2 is favored mainly by the isolated site above the electronic effects, since in both cases, the union of the OOH* is the same. In turn, Ferrara et al. [286] studied the effect of pyrolysis temperature against the selectivity of H2O2 in Co-encapsulated nanoparticles on a nitrogen-doped carbon matrix, determining that higher temperatures (>800 °C) lead to greater H2O2 selectivity, attributed to a larger degree of graphitization and pyrrolic N content. However, the N and O content progressively decreases with increasing pyrolysis temperature, leading to a lower ORR current density.
However, some of the problems of catalysts doped with nonprecious metals are related to their instability in selectivity toward H2O2 with pH. At acidic pHs, the 2e route is favored; however, at basic pHs, the 4e route is given. Therefore, the use of metal oxides such as magnetite has been proposed to improve the ORR 2e activity of catalysts based on carbonaceous materials [287].
With the above, and considering the other limitations associated with both the ORR 2e and Fenton reactions, the doping of carbonaceous materials with structures of the spinel ferrite type is proposed as a viable solution, where the desired bifunctionality is achieved, both for obtaining H2O2 and its subsequent decomposition (Fenton) to OH. This fact, together with the metal active site control and the spontaneous radical OH production in controlled N-doped carbons, opens the gate to the design of new bifunctional catalysts for the simultaneous production of H2O2 and its conversion to OH radicals for the electro-Fenton oxidation of organic pollutants in the water.

6. Bifunctional Electro-Fenton Catalyst for Direct OH Formation

Despite the advances achieved in EF and ORR for the on-site production of H2O2, there are still some problems to overcome such as the secondary pollution related to the Fenton catalysts [288,289] and the requirement of two catalysts: one selective to the reduction of oxygen to H2O2 and another Fenton-type catalyst for the transformation of H2O2 to hydroxyl radicals. Recent efforts are being made to develop materials with dual functionality for the electroreduction of oxygen to H2O2 and Fenton catalysis. However, the preparation of heterogeneous EF catalysts with high selectivity and activity toward ORR via the two-electron pathway is challenging since the active sites for Fenton catalysis are mainly transition metals that usually catalyze oxygen reduction via the 4e route without the production of H2O2.
One strategy to develop bifunctional catalysts capable of generating and activating H2O2 is the doping of carbon materials with different heteroatoms or carbon structures. Li et al. [290] demonstrated that heteroatom-doped carbon materials can convert H2O2 to OH radicals without the need for a Fenton metal catalyst. For that, the authors synthesized an O, F-codoped carbon bifunctional catalyst by the carbonization of polyvinylidene fluoride and proposed that the H2O2 generation depends principally on the ratio C-O/C=O (optimal value in 4), while the activation to OH occurs on the semi-ionic C-F bonds obtaining a sulfamerazine degradation of 90.1% in 3 h during the metal-free EF process. A similar fact was observed by Yang et al. [291] using N-doped graphene. They found that the bifunctional effect is attributed to the presence of N-graphite, which influences the H2O2 generation and N-pyrrolic responsible for H2O2 activation. Similarly, Haider et al. [292] synthesized, in situ, an electrode from polyaniline-derived N-doped carbon nanofibers with bifunctional activity in EF, which was attributed to the content of C=C, oxygen groups, and N-graphitic, which enhanced the H2O2 generation and which can be further activated by N-pyridinic to generate the OH radicals. In turn, Qin et al. [293] used it as a bifunctional metal-free cathode based on O-doped carbon nanotubes. They attributed the excellent metal-free EF behavior of this catalyst to the defects and the C-sp3 that enhances the oxygen adsorption promoting the H2O2 production and to the -C=O active sites that are associated with the OH production. Chen et al. [294] found that groups BC3 and BC2O present in a boride-activated carbon enhanced the oxygen adsorption and facilitated the desorption of OOH, improving the H2O2 generation but also promoting the in situ conversion to OH. Xie et al. [295] showed that the reduced graphene oxide could improve the H2O2 generation, while the MOF (MIL-88A) and carbon felt (MIL-88A/CF) could activate the H2O2 to OH; thus the cathode rGO/MIL-88A/CF resulted in an excellent bifunctional catalyst.
The introduction of different metals in the carbon matrix for use as bifunctional catalysts has also been studied from simple immobilization to more complex structures such as core-shell. Zhang et al. [296] immobilized Fe3O4 in a gas diffusion electrode (GDE) and used it as a rotating cathode for the EF removal of tetracycline. They concluded that the rotating disk gas diffuser greatly enhanced the mass transfer providing a high efficiency in the H2O2 generation, whereas the Fe3O4 active phase was capable of dissociating the H2O2 to OH. Cui et al. [297] also developed a Fe3O4/MWCNTs/CB (carbon black) GDE, which has high performance in the production and activation of H2O2. The presence of CB enhanced the H2O2 production, whereas its activation to OH was ascribed to ≡Fe(II) on the surface of Fe3O4/MWCNTs. Moreover, an improvement in the regeneration of Fe2+ was observed due to the high-speed charge channel of MWCNT (Figure 14). On the other hand, Hu et al. [298] synthesized a bifunctional catalyst based on Fe3C and FeN nanoparticles encapsulated by porous graphitic layers and determined by density function calculations that the active sites for H2O2 generation were the Fe3C, and the activation to OH depend on FeNx sites. Cao et al. [299] designed and synthesized FeOx nanoparticles embedded into N-doped hierarchically porous carbon for its use as a bifunctional catalyst against phenol, sulfamethoxazole, atrazine, rhodamine B, and 2,4-dichlorophenol in neutral water solution. They found that the presence of N pyridinc and N-pyrrolic increased the H2O2 selectivity, and the FeOx nanoparticles could enhance the electron transfer to improve the OH production. Moreover, the Fe3+/Fe2+ cycle could be enhanced by the strong interaction between N-doped porous carbon and FeOx NPs.
Another proposal to improve the production and activation of H2O2 has been the use of catalysts with the presence of two or more different metal or metallic species that can moderate the activity of alone metals and thereby obtain better bifunctional activity. Ghasemi et al. [300] prepared a cathode with carbon nanotubes (CNTs) and CuFe nano-layered double hydroxide showing that the incorporation of CNTs improves the H2O2 generation, and the atoms of Cu and Fe were responsible for the generation of OH. Cui et al. [301] used Cu/CuFe2O4 integrated graphite felt as a bifunctional cathode and attributed the H2O2 generation to Cu0 and the OH production to primary Fe2+ ions. Moreover, the Fe3+/Fe2+ redox cycles could be accelerated by the electron donation of Cu0 and Fe0, and even Cu+. Similarly, Luo et al. [302] degraded 98.1% of tetracycline in 2 h using Cu-doped Fe@Fe2O3 core-shell nanoparticles loaded on nickel foam as a cathode, and determined that Cu0 and Fe0 react with O2 producing H2O2; Fe2+ and Cu+ are responsible for the OH generation from the in situ generated H2O2. Moreover, they proposed that Fe2+ and Cu2+ additional EF active sites could be generated from the redox process between Cu+ and Fe3+ and the replacement reaction between Cu2+ and Fe0. Sun et al. [303] synthesized a nickel foam cathode co-modified with core-shell CoFe alloy/N-doped carbon (CoFe@NC) and carbon nanotubes (CNTs), where the production of H2O2 was attributed to CNTs and the activation to OH to CoFe@NC through Fe2+ and Co2+ oxidation to Fe3+ and Co3+, which are again reduced by Co0/Fe0 and the electric field. Similarly, Yu et al. [304] determined that the carbon felt was the ORR catalyst selective to H2O2, and the Co and Fe atoms generated the OH (Co > Fe) in the cathode of CoFe-layered double hydroxide/carbon felt with bifunctional activity. Li et al. [305] prepared an electrode doped with Fe and Mn oxides coated with a carbon layer and determined that the carbon material was responsible for H2O2 generation, and the metals for OH production. It was also noted that the presence of Mn considerably improves the regeneration of Fe3+/Fe2+ due to Mn that could transfer an electron to Fe to accelerate the cycle.
Bifunctional catalysts using a 3-electron ORR pathway have also been recently proposed as an improved alternative. Xiao et al. [306] reported a catalyst capable of directly generating OH via a 3-electron pathway with FeCo alloy encapsulated by carbon aerogel. This performance was principally associated with the generation of H2O2 by the 2 electron ORR of the COOH group present in the graphite shell, followed by the reduction by one electron toward OH. This 3-electron pathway can be regulated by the electrons from the encapsulated metals.

7. Conclusions

This paper summarizes the different strategies for the design of catalysts with high performance toward Fenton processes and the ORR by the 2e route as a mechanism to improve the efficiency of the electro-Fenton process, finally focusing on the design of bifunctional electro-Fenton catalysts for the production and activation of H2O2.
The key to obtaining high efficiency of ORR 2e is the proper use of structures that allow the end-on adsorption of O2, oxygenated functional groups on the surface of the catalyst, as well as the inclusion of doping atoms that allow synergistic effects toward the selectivity of H2O2. Although the strategies developed have been based on the amount of H2O2 produced and the ORR pathway, some research has opened a gate to the possibility that the morphology of the surface of the catalyst has a much greater impact than expected. Therefore, it is necessary to evaluate, in detail, not only the chemical composition but also other factors such as hydrophobicity, roughness, and/or porosity of the catalyst surface. In general, there are several parameters to consider in the search for an ideal catalyst for ORR 2e, but the amount of N-pyridinic, hydrophobicity, roughness, and doping metal, are so far the most crucial.
As Fenton catalysts, different Fenton-type metals and bimetallic-based materials have been reviewed, raising spinel ferrites as an alternative of interest to provide the bifunctionality for carbonaceous materials, by improving selectivity toward ORR 2e, maintaining the metal ions necessary for Fenton-type catalytic activity suitable for generating OH obtained from the H2O2 generated in the same structure of the catalyst.
Finally, bifunctional electro-Fenton catalysts are described for the production and activation of H2O2. The main strategies are focused on the use of heteroatom-doped carbon materials or carbon-encapsulated Fenton-type metals/carbon composites that moderate the metal’s activity and provide an active site for H2O2 generation by the ORR 2e route, presenting metal sites for the activation of in situ generated H2O2 to ·OH radicals required for the degradation of organic pollutants.

Author Contributions

Conceptualization, A.F.P.-C., F.C.-M. and E.B.-G.; methodology, A.F.P.-C., F.C.-M., E.F.-P. and E.B.-G.; investigation E.F.-P. and A.E.; writing—original draft preparation, E.F.-P.; writing—review and editing, A.E., E.F.-P. and E.B.-G.; supervision, A.F.P.-C. and E.B.-G.; project administration, F.C.-M. and E.B.-G.; funding acquisition, A.F.P.-C. and E.B.-G. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by MICINN grant number [PID2021-127803OB-I00] and Junta de Andalucía [B.RNM.566.UGR20].

Data Availability Statement

Data sharing not applicable.

Acknowledgments

This research has been supported by the Spanish projects from the MICINN (PID2021-127803OB-I00) and Junta de Andalucía (B.RNM.566.UGR20). The authors also thank the “Unidad de Excelencia Química Aplicada a Biomedicina y Medioambiente” of the University of Granada (UEQ—UGR) for its technical assistance. E. Bailón-García is grateful to MICINN for her postdoctoral fellowship (RYC2020-029301-I). E. Fajardo Puerto is grateful to the Colombian Ministry of Sciences, Technology, and Innovation (MINCIENCIAS) for supporting his PhD studies.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Li, L.; Zhao, X.; Liu, D.; Song, K.; Liu, Q.; He, Y. Occurrence and ecological risk assessment of PPCPs in typical inflow rivers of Taihu lake, China. J. Environ. Manag. 2021, 285, 112176. [Google Scholar] [CrossRef]
  2. Archana, G.; Dhodapkar, R.; Kumar, A. Ecotoxicological risk assessment and seasonal variation of some pharmaceuticals and personal care products in the sewage treatment plant and surface water bodies (lakes). Environ. Monit. Assess. 2017, 189, 446. [Google Scholar] [CrossRef] [PubMed]
  3. Luo, Y.; Guo, W.; Ngo, H.H.; Nghiem, L.D.; Hai, F.I.; Zhang, J.; Liang, S.; Wang, X.C. A review on the occurrence of micropollutants in the aquatic environment and their fate and removal during wastewater treatment. Sci. Total Environ. 2014, 473–474, 619–641. [Google Scholar] [CrossRef] [PubMed]
  4. UNESCO. Emerging Pollutants in Water and Wastewater. 2020. Available online: https://en.unesco.org/emergingpollutantsinwaterandwastewater (accessed on 1 July 2022).
  5. Molina, C.B.; Sanz-Santos, E.; Boukhemkhem, A.; Bedia, J.; Belver, C.; Rodriguez, J.J. Removal of emerging pollutants in aqueous phase by heterogeneous Fenton and photo-Fenton with Fe2O3-TiO2-clay heterostructures. Environ. Sci. Pollut. Res. 2020, 27, 38434–38445. [Google Scholar] [CrossRef] [PubMed]
  6. Mailler, R.; Gasperi, J.; Coquet, Y.; Buleté, A.; Vulliet, E.; Deshayes, S.; Zedek, S.; Mirande-Bret, C.; Eudes, V.; Bressy, A.; et al. Removal of a wide range of emerging pollutants from wastewater treatment plant discharges by micro-grain activated carbon in fluidized bed as tertiary treatment at large pilot scale. Sci. Total Environ. 2016, 542, 983–996. [Google Scholar] [CrossRef] [Green Version]
  7. González-Mira, A.; Torreblanca, A.; Hontoria, F.; Navarro, J.C.; Mañanós, E.; Varó, I. Effects of ibuprofen and carbamazepine on the ion transport system and fatty acid metabolism of temperature conditioned juveniles of Solea senegalensis. Ecotoxicol. Environ. Saf. 2018, 148, 693–701. [Google Scholar] [CrossRef] [PubMed]
  8. Kazakova, J.; Villar-Navarro, M.; Ramos-Payán, M.; Aranda-Merino, N.; Román-Hidalgo, C.; Bello-López, M.Á.; Fernández-Torres, R. Monitoring of pharmaceuticals in aquatic biota (Procambarus clarkii) of the Doñana National Park (Spain). J. Environ. Manag. 2021, 297, 113314. [Google Scholar] [CrossRef]
  9. Majumder, A.; Gupta, B.; Gupta, A.K. Pharmaceutically active compounds in aqueous environment: A status, toxicity and insights of remediation. Environ. Res. 2019, 176, 108542. [Google Scholar] [CrossRef]
  10. Rashid, S.S.; Liu, Y.Q. Comparison of life cycle toxicity assessment methods for municipal wastewater treatment with the inclusion of direct emissions of metals, PPCPs and EDCs. Sci. Total Environ. 2021, 756, 143849. [Google Scholar] [CrossRef]
  11. Chaturvedi, P.; Shukla, P.; Giri, B.S.; Chowdhary, P.; Chandra, R.; Gupta, P.; Pandey, A. Prevalence and hazardous impact of pharmaceutical and personal care products and antibiotics in environment: A review on emerging contaminants. Environ. Res. 2021, 194, 110664. [Google Scholar] [CrossRef]
  12. Rodil, R.; Quintana, J.B.; Concha-Graña, E.; López-Mahía, P.; Muniategui-Lorenzo, S.; Prada-Rodríguez, D. Emerging pollutants in sewage, surface and drinking water in Galicia (NW Spain). Chemosphere 2012, 86, 1040–1049. [Google Scholar] [CrossRef]
  13. An, W.; Duan, L.; Zhang, Y.; Zhou, Y.; Wang, B.; Yu, G. Pollution characterization of pharmaceutically active compounds (PhACs) in the northwest of Tai Lake Basin, China: Occurrence, temporal changes, riverine flux and risk assessment. J. Hazard. Mater. 2022, 422, 126889. [Google Scholar] [CrossRef] [PubMed]
  14. Rodriguez-Mozaz, S.; Vaz-Moreira, I.; Varela Della Giustina, S.; Llorca, M.; Barceló, D.; Schubert, S.; Berendonk, T.U.; Michael-Kordatou, I.; Fatta-Kassinos, D.; Martinez, J.L.; et al. Antibiotic residues in final effluents of European wastewater treatment plants and their impact on the aquatic environment. Environ. Int. 2020, 140, 105733. [Google Scholar] [CrossRef]
  15. Kümmerer, K. Antibiotics in the aquatic environment—A review—Part I. Chemosphere 2009, 75, 417–434. [Google Scholar] [CrossRef]
  16. Gelband, H.; Miller-Petrie, M.; Pant, S.; Gandra, S.; Levinson, J.; Barter, D.; White, A.; Laxminarayan, R. The State of the World’s Antibiotics 2015. In Justice A Distance; Center for Disease Dynamics, Economics & Policy: Washington, DC, USA, 2015; pp. 1–30. [Google Scholar] [CrossRef]
  17. Bengtsson-Palme, J.; Kristiansson, E.; Larsson, D.G.J. Environmental factors influencing the development and spread of antibiotic resistance. FEMS Microbiol. Rev. 2018, 42, 68–80. [Google Scholar] [CrossRef] [PubMed]
  18. Dong, J.; Yan, D.; Mo, K.; Chen, Q.; Zhang, J.; Chen, Y.; Wang, Z. Antibiotics along an alpine river and in the receiving lake with a catchment dominated by grazing husbandry. J. Environ. Sci. 2022, 115, 374–382. [Google Scholar] [CrossRef] [PubMed]
  19. Wilke, C. Many of the World’s Rivers are Flush with Dangerous Levels of Antibiotics 2019. Science News, 14 June 2019. [Google Scholar]
  20. Gros, M.; Catalán, N.; Mas-Pla, J.; Čelić, M.; Petrović, M.; Farré, M.J. Groundwater antibiotic pollution and its relationship with dissolved organic matter: Identification and environmental implications. Environ. Pollut. 2021, 289, 117927. [Google Scholar] [CrossRef]
  21. Wang, K.; Zhuang, T.; Su, Z.; Chi, M.; Wang, H. Antibiotic residues in wastewaters from sewage treatment plants and pharmaceutical industries: Occurrence, removal and environmental impacts. Sci. Total Environ. 2021, 788, 147811. [Google Scholar] [CrossRef]
  22. Qian, M.; Yang, L.; Chen, X.; Li, K.; Xue, W.; Li, Y.; Zhao, H.; Cao, G.; Guan, X.; Shen, G. The treatment of veterinary antibiotics in swine wastewater by biodegradation and Fenton-like oxidation. Sci. Total Environ. 2020, 710, 136299. [Google Scholar] [CrossRef]
  23. Jafarisani, M.; Cheshme Khavar, A.H.; Mahjoub, A.R.; Luque, R.; Rodríguez-Padrón, D.; Satari, M.; Gharravi, A.M.; Khastar, H.; Kazemi, S.S.; Masoumikarimi, M. Enhanced visible-light-driven photocatalytic degradation of emerging water contaminants by a modified zinc oxide-based photocatalyst; In-vivo and in-vitro toxicity evaluation of wastewater and PCO-treated water. Sep. Purif. Technol. 2020, 243, 116430. [Google Scholar] [CrossRef]
  24. Ata, R.; Yıldız Töre, G. Characterization and removal of antibiotic residues by NFC-doped photocatalytic oxidation from domestic and industrial secondary treated wastewaters in Meric-Ergene Basin and reuse assessment for irrigation. J. Environ. Manag. 2019, 233, 673–680. [Google Scholar] [CrossRef] [PubMed]
  25. Zha, S.; Cheng, Y.; Gao, Y.; Chen, Z.; Megharaj, M.; Naidu, R. Nanoscale zero-valent iron as a catalyst for heterogeneous Fenton oxidation of amoxicillin. Chem. Eng. J. 2014, 255, 141–148. [Google Scholar] [CrossRef]
  26. Wang, A.; Zhang, Y.; Han, S.; Guo, C.; Wen, Z.; Tian, X.; Li, J. Electro-Fenton oxidation of a β-lactam antibiotic cefoperazone: Mineralization, biodegradability and degradation mechanism. Chemosphere 2021, 270, 129486. [Google Scholar] [CrossRef] [PubMed]
  27. Wang, C.; Lin, C.Y.; Liao, G.Y. Degradation of antibiotic tetracycline by ultrafine-bubble ozonation process. J. Water Process Eng. 2020, 37, 101463. [Google Scholar] [CrossRef]
  28. Baaloudj, O.; Assadi, I.; Nasrallah, N.; El Jery, A.; Khezami, L.; Assadi, A.A. Simultaneous removal of antibiotics and inactivation of antibiotic-resistant bacteria by photocatalysis: A review. J. Water Process Eng. 2021, 42, 102089. [Google Scholar] [CrossRef]
  29. Baran, W.; Adamek, E.; Jajko, M.; Sobczak, A. Removal of veterinary antibiotics from wastewater by electrocoagulation. Chemosphere 2018, 194, 381–389. [Google Scholar] [CrossRef]
  30. Salas, G. Tratamiento por Oxidación Avanzada (Reacción Fenton) de Aguas Residuales de la Industria Textil. Rev. Peru. Quim. Ing. Quim. 2010, 12, 30–38. [Google Scholar]
  31. Fenton, H.J. Oxidation of Tartaric Acid in presence of Iron. J. Chem. Soc. Trans. 1894, 65, 899–910. [Google Scholar] [CrossRef] [Green Version]
  32. Huang, Y.; Liu, H.; Liu, S.; Li, C.; Yuan, S. Glucose oxidase modified Fenton reactions for in-situ ROS generation and potential application in groundwater remediation. Chemosphere 2020, 253, 126648. [Google Scholar] [CrossRef]
  33. Ramos, J.M.P.; Pereira-Queiroz, N.M.; Santos, D.H.S.; Nascimento, J.R.; de Carvalho, C.M.; Tonholo, J.; Zanta, C.L.P.S. Printing ink effluent remediation: A comparison between electrochemical and Fenton treatments. J. Water Process Eng. 2019, 31, 100803. [Google Scholar] [CrossRef]
  34. Usman, M.; Ho, Y.S. A bibliometric study of the Fenton oxidation for soil and water remediation. J. Environ. Manag. 2020, 270, 110886. [Google Scholar] [CrossRef] [PubMed]
  35. Wong, K.T.; Jang, S.B.; Choong, C.E.; won Kang, C.; Lee, G.C.; Song, J.Y.; Yoon, Y.; Jang, M. In situ Fenton remediation for diesel contaminated clayey zone assisted by thermal plasma blasting: Synergism and cost estimation. Chemosphere 2022, 286, 131574. [Google Scholar] [CrossRef] [PubMed]
  36. Jain, B.; Singh, A.K.; Kim, H.; Lichtfouse, E.; Sharma, V.K. Treatment of organic pollutants by homogeneous and heterogeneous Fenton reaction processes. Environ. Chem. Lett. 2018, 16, 947–967. [Google Scholar] [CrossRef] [Green Version]
  37. Sharma, M.; Mohapatra, T.; Ghosh, P. Hydrodynamics, mass and heat transfer study for emerging heterogeneous Fenton process in multiphase fluidized-bed reactor system for wastewater treatment—A review. Chem. Eng. Res. Des. 2021, 171, 48–62. [Google Scholar] [CrossRef]
  38. Yu, G.H.; Kuzyakov, Y. Fenton chemistry and reactive oxygen species in soil: Abiotic mechanisms of biotic processes, controls and consequences for carbon and nutrient cycling. Earth-Sci. Rev. 2021, 214, 103525. [Google Scholar] [CrossRef]
  39. Gligorovski, S.; Strekowski, R.; Barbati, S.; Vione, D. Environmental Implications of Hydroxyl Radicals (OH). Chem. Rev. 2015, 115, 13051–13092. [Google Scholar] [CrossRef] [PubMed]
  40. Amir, J.; Rojas, A.; Fernando, L.; Giraldo, G.; Trujillo, J.M. Empleo del reactivo de Fenton para la degradación del colorante Tartrazina. Lasallista Investig. 2009, 6, 27–34. [Google Scholar]
  41. Zhang, X.; Zhang, Y.; Yu, Z.; Wei, X.; Wu, W.D.; Wang, X.; Wu, Z. Facile synthesis of mesoporous anatase/rutile/hematite triple heterojunctions for superior heterogeneous photo-Fenton catalysis. Appl. Catal. B Environ. 2020, 263, 118335. [Google Scholar] [CrossRef]
  42. Song, W.; Cheng, M.; Ma, J.; Ma, W.; Chen, C.; Zhao, J. Decomposition of hydrogen peroxide driven by photochemical cycling of iron species in clay. Environ. Sci. Technol. 2006, 40, 4782–4787. [Google Scholar] [CrossRef]
  43. Wang, N.; Zheng, T.; Zhang, G.; Wang, P. A review on Fenton-like processes for organic wastewater treatment. J. Environ. Chem. Eng. 2016, 4, 762–787. [Google Scholar] [CrossRef] [Green Version]
  44. Ochando-Pulido, J.M.; Pimentel-Moral, S.; Verardo, V.; Martinez-Ferez, A. A focus on advanced physico-chemical processes for olive mill wastewater treatment. Sep. Purif. Technol. 2017, 179, 161–174. [Google Scholar] [CrossRef]
  45. Babuponnusami, A.; Muthukumar, K. A review on Fenton and improvements to the Fenton process for wastewater treatment. J. Environ. Chem. Eng. 2014, 2, 557–572. [Google Scholar] [CrossRef]
  46. Bokare, A.D.; Choi, W. Review of iron-free Fenton-like systems for activating H2O2 in advanced oxidation processes. J. Hazard. Mater. 2014, 275, 121–135. [Google Scholar] [CrossRef] [PubMed]
  47. Fernández, L.; González-Rodríguez, J.; Gamallo, M.; Vargas-Osorio, Z.; Vázquez-Vázquez, C.; Piñeiro, Y.; Rivas, J.; Feijoo, G.; Moreira, M.T. Iron oxide-mediated photo-Fenton catalysis in the inactivation of enteric bacteria present in wastewater effluents at neutral pH. Environ. Pollut. 2020, 266, 115181. [Google Scholar] [CrossRef] [PubMed]
  48. Wang, L.; Zhang, Y.; Qian, J. Graphene aerogel-based catalysts in Fenton-like reactions for water decontamination: A short review. Chem. Eng. J. Adv. 2021, 8, 100171. [Google Scholar] [CrossRef]
  49. Dinçer, A.R.; Çifçi, D.İ.; Cinkaya, D.D.; Dülger, E.; Karaca, F. Treatment of organic peroxide containing wastewater and water recovery by fenton-adsorption and fenton-nanofiltration processes. J. Environ. Manag. 2021, 299, 113557. [Google Scholar] [CrossRef] [PubMed]
  50. Hu, C.; Huang, D.; Zeng, G.; Cheng, M.; Gong, X.; Wang, R.; Xue, W.; Hu, Z.; Liu, Y. The combination of Fenton process and Phanerochaete chrysosporium for the removal of bisphenol A in river sediments: Mechanism related to extracellular enzyme, organic acid and iron. Chem. Eng. J. 2018, 338, 432–439. [Google Scholar] [CrossRef]
  51. Soufi, A.; Hajjaoui, H.; Elmoubarki, R.; Abdennouri, M.; Qourzal, S.; Barka, N. Spinel ferrites nanoparticles: Synthesis methods and application in heterogeneous Fenton oxidation of organic pollutants—A review. Appl. Surf. Sci. Adv. 2021, 6, 100145. [Google Scholar] [CrossRef]
  52. Cheng, M.; Zeng, G.; Huang, D.; Lai, C.; Xu, P.; Zhang, C.; Liu, Y. Hydroxyl radicals based advanced oxidation processes (AOPs) for remediation of soils contaminated with organic compounds: A review. Chem. Eng. J. 2016, 284, 582–598. [Google Scholar] [CrossRef]
  53. Rao, S.N.; Shrivastava, S. Treatment of complex recalcitrant wastewater using Fenton process. Mater. Today Proc. 2020, 29, 1161–1165. [Google Scholar] [CrossRef]
  54. Zhao, X.; Chen, L.; Ma, H.; Ma, J.; Gao, D. Effective removal of polymer quaternary ammonium salt by biodegradation and a subsequent Fenton oxidation process. Ecotoxicol. Environ. Saf. 2020, 188, 109919. [Google Scholar] [CrossRef] [PubMed]
  55. Liu, C.; Yi, H.; Yang, B.; Jia, F.; Song, S. Activation of Fenton reaction by controllable oxygen incorporation in MoS2-Fe under visible light irradiation. Appl. Surf. Sci. 2021, 566, 150674. [Google Scholar] [CrossRef]
  56. Qin, H.; Cheng, H.; Li, H.; Wang, Y. Degradation of ofloxacin, amoxicillin and tetracycline antibiotics using magnetic core–shell MnFe2O4@C-NH2 as a heterogeneous Fenton catalyst. Chem. Eng. J. 2020, 396, 125304. [Google Scholar] [CrossRef]
  57. Liu, Y.; Zhang, X.; Deng, J.; Liu, Y. A novel CNTs-Fe3O4 synthetized via a ball-milling strategy as efficient fenton-like catalyst for degradation of sulfonamides. Chemosphere 2021, 277, 130305. [Google Scholar] [CrossRef]
  58. Fornazaria, A.L.; Labriola, V.F.; Da Silva, B.F.; Castro, L.F.; Perussi, J.R.; Vieira, E.M.; Azevedo, E.B. Coupling Zero-Valent Iron and Fenton processes for degrading sulfamethazine, sulfathiazole, and norfloxacin. J. Environ. Chem. Eng. 2021, 9, 105761. [Google Scholar] [CrossRef]
  59. Qin, H.; Yang, Y.; Shi, W.; She, Y.; Wu, S. Heterogeneous Fenton degradation of azithromycin antibiotic in water catalyzed by amino/thiol-functionalized MnFe2O4 magnetic nanocatalysts. J. Environ. Chem. Eng. 2021, 9, 106184. [Google Scholar] [CrossRef]
  60. Li, T.; Chen, Y.; Wang, X.; Liang, J.; Zhou, L. Modifying organic carbon in Fe3O4-loaded schwertmannite to improve heterogeneous Fenton activity through accelerating Fe(II) generation. Appl. Catal. B Environ. 2021, 285, 119830. [Google Scholar] [CrossRef]
  61. Goswami, A.; Jiang, J.Q.; Petri, M. Treatability of five micro-pollutants using modified Fenton reaction catalysed by zero-valent iron powder (Fe(0)). J. Environ. Chem. Eng. 2021, 9, 105393. [Google Scholar] [CrossRef]
  62. Homem, V.; Alves, A.; Santos, L. Amoxicillin degradation at ppb levels by Fenton’s oxidation using design of experiments. Sci. Total Environ. 2010, 408, 6272–6280. [Google Scholar] [CrossRef]
  63. Cheng, S.; Zhang, C.; Li, J.; Pan, X.; Zhai, X.; Jiao, Y.; Li, Y.; Dong, W.; Qi, X. Highly efficient removal of antibiotic from biomedical wastewater using Fenton-like catalyst magnetic pullulan hydrogels. Carbohydr. Polym. 2021, 262, 117951. [Google Scholar] [CrossRef]
  64. Wang, C.; Sun, R.; Huang, R.; Wang, H. Superior fenton-like degradation of tetracycline by iron loaded graphitic carbon derived from microplastics: Synthesis, catalytic performance, and mechanism. Sep. Purif. Technol. 2021, 270, 118773. [Google Scholar] [CrossRef]
  65. Del Álamo, A.C.; González, C.; Pariente, M.I.; Molina, R.; Martínez, F. Fenton-like catalyst based on a reticulated porous perovskite material: Activity and stability for the on-site removal of pharmaceutical micropollutans in a hospital wastewater. Chem. Eng. J. 2020, 401, 126113. [Google Scholar] [CrossRef]
  66. Wang, C.; Yu, G.; Chen, H.; Wang, J. Degradation of norfloxacin by hydroxylamine enhanced fenton system: Kinetics, mechanism and degradation pathway. Chemosphere 2021, 270, 129408. [Google Scholar] [CrossRef]
  67. Salari, M.; Rakhshandehroo, G.R.; Nikoo, M.R.; Zerafat, M.M.; Mooselu, M.G. Optimal degradation of Ciprofloxacin in a heterogeneous Fenton-like process using (δ-FeOOH)/MWCNTs nanocomposite. Environ. Technol. Innov. 2021, 23, 101625. [Google Scholar] [CrossRef]
  68. Karimnezhad, H.; Navarchian, A.H.; Tavakoli Gheinani, T.; Zinadini, S. Amoxicillin removal by Fe-based nanoparticles immobilized on polyacrylonitrile membrane: Individual nanofiltration or Fenton reaction, vs. engineered combined process. Chem. Eng. Res. Des. 2020, 153, 187–200. [Google Scholar] [CrossRef]
  69. Chen, P.; Sun, F.; Wang, W.; Tan, F.; Wang, X.; Qiao, X. Facile one-pot fabrication of ZnO2 particles for the efficient Fenton-like degradation of tetracycline. J. Alloys Compd. 2020, 834, 155220. [Google Scholar] [CrossRef]
  70. Yang, W.; Hong, P.; Yang, D.; Yang, Y.; Wu, Z.; Xie, C.; He, J.; Zhang, K.; Kong, L.; Liu, J. Enhanced Fenton-like degradation of sulfadiazine by single atom iron materials fixed on nitrogen-doped porous carbon. J. Colloid Interface Sci. 2021, 597, 56–65. [Google Scholar] [CrossRef]
  71. de Melo Costa-Serge, N.; Gonçalves, R.G.L.; Rojas-Mantilla, H.D.; Santilli, C.V.; Hammer, P.; Nogueira, R.F.P. Fenton-like degradation of sulfathiazole using copper-modified MgFe-CO3 layered double hydroxide. J. Hazard. Mater. 2021, 413, 125388. [Google Scholar] [CrossRef]
  72. Liu, L.; Xu, Q.; Owens, G.; Chen, Z. Fenton-oxidation of rifampicin via a green synthesized rGO@nFe/Pd nanocomposite. J. Hazard. Mater. 2021, 402, 123544. [Google Scholar] [CrossRef]
  73. Yi, Y.; Luo, J.; Fang, Z. Magnetic biochar derived from Eichhornia crassipes for highly efficient Fenton-like degradation of antibiotics: Mechanism and contributions. J. Environ. Chem. Eng. 2021, 9, 106258. [Google Scholar] [CrossRef]
  74. Gao, Y.; Zhu, W.; Liu, J.; Lin, P.; Zhang, J.; Huang, T.; Liu, K. Mesoporous sulfur-doped CoFe2O4 as a new Fenton catalyst for the highly efficient pollutants removal. Appl. Catal. B Environ. 2021, 295, 120273. [Google Scholar] [CrossRef]
  75. Carrasco-Díaz, M.R.; Castillejos-López, E.; Cerpa-Naranjo, A.; Rojas-Cervantes, M.L. Efficient removal of paracetamol using LaCu1−xMxO3 (M = Mn, Ti) perovskites as heterogeneous Fenton-like catalysts. Chem. Eng. J. 2016, 304, 408–418. [Google Scholar] [CrossRef]
  76. Guo, H.; Li, Z.; Xie, Z.; Song, J.; Xiang, L.; Zhou, L.; Li, C.; Liao, L.; Li, J.; Wang, H. Accelerated Fenton reaction for antibiotic ofloxacin degradation in discharge plasma system based on graphene-Fe3O4 nanocomposites. Vacuum 2021, 185, 110022. [Google Scholar] [CrossRef]
  77. Wang, X.; Jin, H.; Wu, D.; Nie, Y.; Tian, X.; Yang, C.; Zhou, Z.; Li, Y. Fe3O4@S-doped ZnO: A magnetic, recoverable, and reusable Fenton-like catalyst for efficient degradation of ofloxacin under alkaline conditions. Environ. Res. 2020, 186, 109626. [Google Scholar] [CrossRef]
  78. Sun, F.; Liu, H.; Wang, H.; Shu, D.; Chen, T.; Zou, X.; Huang, F.; Chen, D. A novel discovery of a heterogeneous Fenton-like system based on natural siderite: A wide range of pH values from 3 to 9. Sci. Total Environ. 2020, 698, 134293. [Google Scholar] [CrossRef]
  79. Ma, E.; Wang, K.; Hu, Z.; Wang, H. Dual-stimuli-responsive CuS-based micromotors for efficient photo-Fenton degradation of antibiotics. J. Colloid Interface Sci. 2021, 603, 685–694. [Google Scholar] [CrossRef]
  80. Gu, W.; Huang, X.; Tian, Y.; Cao, M.; Zhou, L.; Zhou, Y.; Lu, J.; Lei, J.; Zhou, Y.; Wang, L.; et al. High-efficiency adsorption of tetracycline by cooperation of carbon and iron in a magnetic Fe/porous carbon hybrid with effective Fenton regeneration. Appl. Surf. Sci. 2021, 538, 147813. [Google Scholar] [CrossRef]
  81. Liao, Q.; Wang, D.; Ke, C.; Zhang, Y.; Han, Q.; Zhang, Y.; Xi, K. Metal-free Fenton-like photocatalysts based on covalent organic frameworks. Appl. Catal. B Environ. 2021, 298, 120548. [Google Scholar] [CrossRef]
  82. Luo, L.; Wang, G.; Wang, Z.; Ma, J.; He, Y.; He, J.; Wang, L.; Liu, Y.; Xiao, H.; Xiao, Y.; et al. Optimization of Fenton process on removing antibiotic resistance genes from excess sludge by single-factor experiment and response surface methodology. Sci. Total Environ. 2021, 788, 147889. [Google Scholar] [CrossRef]
  83. Teel, A.L.; Warberg, C.R.; Atkinson, D.A.; Watts, R.J. Comparison of mineral and soluble iron Fenton’s catalysts for the treatment of trichloroethylene. Water Res. 2001, 35, 977–984. [Google Scholar] [CrossRef]
  84. Lewis, S.; Lynch, A.; Bachas, L.; Hampson, S.; Ormsbee, L.; Bhattacharyya, D. Chelate-modified fenton reaction for the degradation of trichloroethylene in aqueous and two-phase systems. Environ. Eng. Sci. 2009, 26, 849–859. [Google Scholar] [CrossRef]
  85. Bello, M.M.; Abdul Raman, A.A.; Asghar, A. A review on approaches for addressing the limitations of Fenton oxidation for recalcitrant wastewater treatment. Process Saf. Environ. Prot. 2019, 126, 119–140. [Google Scholar] [CrossRef]
  86. Huang, C.; Zhang, C.; Huang, D.; Wang, D.; Tian, S.; Wang, R.; Yang, Y.; Wang, W.; Qin, F. Influence of surface functionalities of pyrogenic carbonaceous materials on the generation of reactive species towards organic contaminants: A review. Chem. Eng. J. 2021, 404, 127066. [Google Scholar] [CrossRef]
  87. Dumesic, J.A.; Huber, G.W.; Boudart, M. Principles of Heterogeneous Catalysis. In Handbook of Heterogeneous Catalysis; Wiley: Hoboken, NJ, USA, 2008. [Google Scholar] [CrossRef]
  88. Expósito, E.; Sánchez-Sánchez, C.M.; Montiel, V. Mineral Iron Oxides as Iron Source in Electro-Fenton and Photoelectro-Fenton Mineralization Processes. J. Electrochem. Soc. 2007, 154, E116. [Google Scholar] [CrossRef]
  89. Bai, Z.; Yang, Q.; Wang, J. Degradation of sulfamethazine antibiotics in Fenton-like system using Fe3O4 magnetic nanoparticles as catalyst. Environ. Prog. Sustain. Energy 2017, 36, 1743–1753. [Google Scholar] [CrossRef]
  90. Xie, W.; Huang, Z.; Zhou, F.; Li, Y.; Bi, X.; Bian, Q.; Sun, S. Heterogeneous fenton-like degradation of amoxicillin using MOF-derived Fe0 embedded in mesoporous carbon as an effective catalyst. J. Clean. Prod. 2021, 313, 127754. [Google Scholar] [CrossRef]
  91. Furia, F.; Minella, M.; Gosetti, F.; Turci, F.; Sabatino, R.; Di Cesare, A.; Corno, G.; Vione, D. Elimination from wastewater of antibiotics reserved for hospital settings, with a Fenton process based on zero-valent iron. Chemosphere 2021, 283, 131170. [Google Scholar] [CrossRef]
  92. Rodriguez-Rodriguez, J.; Ochando-Pulido, J.M.; Martinez-Ferez, A. The effect of pH in tannery wastewater by Fenton vs. Heterogeneous Fenton process. Chem. Eng. Trans. 2019, 73, 205–210. [Google Scholar] [CrossRef]
  93. He, J.; Yang, X.; Men, B.; Wang, D. Interfacial mechanisms of heterogeneous Fenton reactions catalyzed by iron-based materials: A review. J. Environ. Sci. 2016, 39, 97–109. [Google Scholar] [CrossRef]
  94. Liu, G.; Zhang, Y.; Yu, H.; Jin, R.; Zhou, J. Acceleration of goethite-catalyzed Fenton-like oxidation of ofloxacin by biochar. J. Hazard. Mater. 2020, 397, 122783. [Google Scholar] [CrossRef]
  95. Huang, X.; Hou, X.; Zhao, J.; Zhang, L. Hematite facet confined ferrous ions as high efficient Fenton catalysts to degrade organic contaminants by lowering H2O2 decomposition energetic span. Appl. Catal. B Environ. 2016, 181, 127–137. [Google Scholar] [CrossRef]
  96. Yang, S.; Xiong, Y.; Ge, Y.; Zhang, S. Heterogeneous Fenton oxidation of nitric oxide by magnetite: Kinetics and mechanism. Mater. Lett. 2018, 218, 257–261. [Google Scholar] [CrossRef]
  97. Liu, W.; Zhou, J.; Liu, D.; Liu, S.; Liu, X.; Xiao, S. Enhancing electronic transfer by magnetic iron materials and metal-organic framework via heterogeneous Fenton-like process and photocatalysis. Mater. Sci. Semicond. Process 2021, 135, 106096. [Google Scholar] [CrossRef]
  98. Hassani, A.; Karaca, M.; Karaca, S.; Khataee, A.; Açışlı, Ö.; Yılmaz, B. Preparation of magnetite nanoparticles by high-energy planetary ball mill and its application for ciprofloxacin degradation through heterogeneous Fenton process. J. Environ. Manag. 2018, 211, 53–62. [Google Scholar] [CrossRef]
  99. Gholizadeh, A.M.; Zarei, M.; Ebratkhahan, M.; Hasanzadeh, A. Phenazopyridine degradation by electro-Fenton process with magnetite nanoparticles-activated carbon cathode, artificial neural networks modeling. J. Environ. Chem. Eng. 2021, 9, 104999. [Google Scholar] [CrossRef]
  100. Gonçalves, R.G.L.; Lopes, P.A.; Resende, J.A.; Pinto, F.G.; Tronto, J.; Guerreiro, M.C.; de Oliveira, L.C.A.; de Castro Nunes, W.; Neto, J.L. Performance of magnetite/layered double hydroxide composite for dye removal via adsorption, Fenton and photo-Fenton processes. Appl. Clay Sci. 2019, 179, 105152. [Google Scholar] [CrossRef]
  101. Rial, J.B.; Ferreira, M.L. Challenges of dye removal treatments based on IONzymes: Beyond heterogeneous Fenton. J. Water Process Eng. 2021, 41, 102065. [Google Scholar] [CrossRef]
  102. Yang, Z.; Yu, A.; Shan, C.; Gao, G.; Pan, B. Enhanced Fe(III)-mediated Fenton oxidation of atrazine in the presence of functionalized multi-walled carbon nanotubes. Water Res. 2018, 137, 37–46. [Google Scholar] [CrossRef]
  103. Seo, J.; Lee, H.J.; Lee, H.; Kim, H.E.; Lee, J.Y.; Kim, H.S.; Lee, C. Enhanced production of reactive oxidants by Fenton-like reactions in the presence of carbon materials. Chem. Eng. J. 2015, 273, 502–508. [Google Scholar] [CrossRef]
  104. Qin, Y.; Zhang, L.; An, T. Hydrothermal Carbon-Mediated Fenton-like Reaction Mechanism in the Degradation of Alachlor: Direct Electron Transfer from Hydrothermal Carbon to Fe(III). ACS Appl. Mater. Interfaces 2017, 9, 17115–17124. [Google Scholar] [CrossRef]
  105. Bach, A.; Semiat, R. The role of activated carbon as a catalyst in GAC/iron oxide/H2O2 oxidation process. Desalination 2011, 273, 57–63. [Google Scholar] [CrossRef]
  106. Kappler, A.; Wuestner, M.L.; Ruecker, A.; Harter, J.; Halama, M.; Behrens, S. Biochar as an Electron Shuttle between Bacteria and Fe(III) Minerals. Environ. Sci. Technol. Lett. 2014, 1, 339–344. [Google Scholar] [CrossRef]
  107. Sun, L.; Yao, Y.; Wang, L.; Mao, Y.; Huang, Z.; Yao, D.; Lu, W.; Chen, W. Efficient removal of dyes using activated carbon fibers coupled with 8-hydroxyquinoline ferric as a reusable Fenton-like catalyst. Chem. Eng. J. 2014, 240, 413–419. [Google Scholar] [CrossRef]
  108. Yao, T.; Qi, Y.; Mei, Y.; Yang, Y.; Aleisa, R.; Tong, X.; Wu, J. One-step preparation of reduced graphene oxide aerogel loaded with mesoporous copper ferrite nanocubes: A highly efficient catalyst in microwave-assisted Fenton reaction. J. Hazard. Mater. 2019, 378, 105. [Google Scholar] [CrossRef]
  109. Kuśmierek, K.; Kiciński, W.; Norek, M.; Polański, M.; Budner, B. Oxidative and adsorptive removal of chlorophenols over Fe-, N- and S-multi-doped carbon xerogels. J. Environ. Chem. Eng. 2021, 9, 105568. [Google Scholar] [CrossRef]
  110. Cruz, D.R.S.; de Jesus, G.K.; Santos, C.A.; Silva, W.R.; Wisniewski, A.; Cunha, G.C.; Romão, L.P.C. Magnetic nanostructured material as heterogeneous catalyst for degradation of AB210 dye in tannery wastewater by electro-Fenton process. Chemosphere 2021, 280, 130675. [Google Scholar] [CrossRef] [PubMed]
  111. Zhang, X.; Guo, Y.; Shi, S.; Liu, E.; Li, T.; Wei, S.; Li, Y.; Li, Y.; Sun, G.; Zhao, Z. Efficient and stable iron-copper montmorillonite heterogeneous Fenton catalyst for removing Rhodamine B. Chem. Phys. Lett. 2021, 776, 138673. [Google Scholar] [CrossRef]
  112. Li, X.; Tian, Z.; Liang, K.; Wang, Y. Enhanced photo-Fenton degradation performance over multi-metal co-supported SAPO-18 zeolites by promoted active species yield. J. Photochem. Photobiol. A Chem. 2019, 379, 79–87. [Google Scholar] [CrossRef]
  113. Xu, X.; Chen, W.; Zong, S.; Ren, X.; Liu, D. Magnetic clay as catalyst applied to organics degradation in a combined adsorption and Fenton-like process. Chem. Eng. J. 2019, 373, 140–149. [Google Scholar] [CrossRef]
  114. Ferroudj, N.; Beaunier, P.; Davidson, A.; Abramson, S. Fe3+-doped ordered mesoporous γ-Fe2O3/SiO2 microspheres as a highly efficient magnetically separable heterogeneous Fenton catalyst. Microporous Mesoporous Mater. 2021, 326, 111373. [Google Scholar] [CrossRef]
  115. Wei, K.; Liu, X.; Cao, S.; Cui, H.; Zhang, Y.; Ai, Z. Fe2O3@FeB composites facilitate heterogeneous Fenton process by efficient Fe(III)/Fe(II) cycle and in-situ H2O2 generation. Chem. Eng. J. Adv. 2021, 8, 100165. [Google Scholar] [CrossRef]
  116. Carrasco-Díaz, M.R.; Castillejos-López, E.; Cerpa-Naranjo, A.; Rojas-Cervantes, M.L. On the textural and crystalline properties of Fe-carbon xerogels. Application as Fenton-like catalysts in the oxidation of paracetamol by H2O2. Microporous Mesoporous Mater. 2017, 237, 282–293. [Google Scholar] [CrossRef]
  117. Asghar, A.; Raman, A.A.A.; Daud, W.M.A.W. Advanced oxidation processes for in-situ production of hydrogen peroxide/hydroxyl radical for textile wastewater treatment: A review. J. Clean. Prod. 2015, 87, 826–838. [Google Scholar] [CrossRef] [Green Version]
  118. Yazici Guvenc, S.; Dincer, K.; Varank, G. Performance of electrocoagulation and electro-Fenton processes for treatment of nanofiltration concentrate of biologically stabilized landfill leachate. J. Water Process Eng. 2019, 31, 100863. [Google Scholar] [CrossRef]
  119. Tian, Y.; Zhou, M.; Pan, Y.; Cai, J.; Ren, G. Pre-magnetized Fe0 as heterogeneous electro-Fenton catalyst for the degradation of p-nitrophenol at neutral pH. Chemosphere 2020, 240, 124962. [Google Scholar] [CrossRef] [PubMed]
  120. Huang, L.Z.; Zhu, M.; Liu, Z.; Wang, Z.; Hansen, H.C.B. Single sheet iron oxide: An efficient heterogeneous electro-Fenton catalyst at neutral pH. J. Hazard. Mater. 2019, 364, 39–47. [Google Scholar] [CrossRef]
  121. Midassi, S.; Bedoui, A.; Bensalah, N. Efficient degradation of chloroquine drug by electro-Fenton oxidation: Effects of operating conditions and degradation mechanism. Chemosphere 2020, 260, 127558. [Google Scholar] [CrossRef]
  122. Elbatea, A.A.; Nosier, S.A.; Zatout, A.A.; Hassan, I.; Sedahmed, G.H.; Abdel-Aziz, M.H.; El-Naggar, M.A. Removal of reactive red 195 from dyeing wastewater using electro-Fenton process in a cell with oxygen sparged fixed bed electrodes. J. Water Process Eng. 2021, 41, 102042. [Google Scholar] [CrossRef]
  123. Senthilnathan, J.; Younis, S.A.; Kwon, E.E.; Surenjan, A.; Kim, K.H.; Yoshimura, M. An efficient system for electro-Fenton oxidation of pesticide by a reduced graphene oxide-aminopyrazine@3DNi foam gas diffusion electrode. J. Hazard. Mater. 2020, 400, 123323. [Google Scholar] [CrossRef]
  124. McQuillan, R.V.; Stevens, G.W.; Mumford, K.A. Assessment of the electro-Fenton pathway for the removal of naphthalene from contaminated waters in remote regions. Sci. Total Environ. 2021, 762, 143155. [Google Scholar] [CrossRef]
  125. Poza-Nogueiras, V.; Rosales, E.; Pazos, M.; Sanromán, M.Á. Current advances and trends in electro-Fenton process using heterogeneous catalysts—A review. Chemosphere 2018, 201, 399–416. [Google Scholar] [CrossRef] [PubMed]
  126. Matyszczak, G.; Krzyczkowska, K.; Fidler, A. A novel, two-electron catalysts for the electro-Fenton process. J. Water Process Eng. 2020, 36, 101242. [Google Scholar] [CrossRef]
  127. Liang, J.; Xiang, Q.; Lei, W.; Zhang, Y.; Sun, J.; Zhu, H.; Wang, S. Ferric iron reduction reaction electro-Fenton with gas diffusion device: A novel strategy for improvement of comprehensive efficiency in electro-Fenton. J. Hazard. Mater. 2021, 412, 125195. [Google Scholar] [CrossRef] [PubMed]
  128. Zhang, C.; Zhou, M.; Yu, X.; Ma, L.; Yu, F. Modified iron-carbon as heterogeneous electro-Fenton catalyst for organic pollutant degradation in near neutral pH condition: Characterization, degradation activity and stability. Electrochim. Acta 2015, 160, 254–262. [Google Scholar] [CrossRef]
  129. Zhu, W.; Li, Y.; Gao, Y.; Wang, C.; Zhang, J.; Bai, H.; Huang, T. A new method to fabricate the cathode by cyclic voltammetric electrodeposition for electro-Fenton application. Electrochim. Acta 2020, 349, 136415. [Google Scholar] [CrossRef]
  130. Ganzenko, O.; Trellu, C.; Oturan, N.; Huguenot, D.; Péchaud, Y.; van Hullebusch, E.D.; Oturan, M.A. Electro-Fenton treatment of a complex pharmaceutical mixture: Mineralization efficiency and biodegradability enhancement. Chemosphere 2020, 253, 126659. [Google Scholar] [CrossRef]
  131. Robles, I.; Becerra, E.; Barrios, J.A.; Maya, C.; Jiménez, B.; Rodríguez-Valadez, F.J.; Rivera, F.; García-Espinoza, J.D.; Godínez, L.A. Inactivation of helminth eggs in an electro-Fenton reactor: Towards full electrochemical disinfection of human waste using activated carbon. Chemosphere 2020, 250, 126260. [Google Scholar] [CrossRef]
  132. Wang, Y.; Liu, Y.; Liu, T.; Song, S.; Gui, X.; Liu, H.; Tsiakaras, P. Dimethyl phthalate degradation at novel and efficient electro-Fenton cathode. Appl. Catal. B Environ. 2014, 156–157, 1–7. [Google Scholar] [CrossRef]
  133. Zhao, K.; Quan, X.; Chen, S.; Yu, H.; Zhang, Y.; Zhao, H. Enhanced electro-Fenton performance by fluorine-doped porous carbon for removal of organic pollutants in wastewater. Chem. Eng. J. 2018, 354, 606–615. [Google Scholar] [CrossRef]
  134. Liu, Y.; Chen, S.; Quan, X.; Yu, H.; Zhao, H.; Zhang, Y. Efficient Mineralization of Perfluorooctanoate by Electro-Fenton with H2O2 Electro-generated on hierarchically Porous Carbon. Environ. Sci. Technol. 2015, 49, 12528–12533. [Google Scholar] [CrossRef]
  135. Sun, C.; Chen, T.; Huang, Q.; Duan, X.; Zhan, M.; Ji, L.; Li, X.; Wang, S.; Yan, J. Biochar cathode: Reinforcing electro-Fenton pathway against four-electron reduction by controlled carbonization and surface chemistry. Sci. Total Environ. 2021, 754, 142136. [Google Scholar] [CrossRef] [PubMed]
  136. Ergan, B.T.; Gengec, E. Dye degradation and kinetics of online Electro-Fenton system with thermally activated carbon fiber cathodes. J. Environ. Chem. Eng. 2020, 8, 104217. [Google Scholar] [CrossRef]
  137. Ramírez-Pereda, B.; Álvarez-Gallegos, A.; Rangel-Peraza, J.G.; Bustos-Terrones, Y.A. Kinetics of Acid Orange 7 oxidation by using carbon fiber and reticulated vitreous carbon in an electro-Fenton process. J. Environ. Manag. 2018, 213, 279–287. [Google Scholar] [CrossRef]
  138. Xia, G.; Lu, Y.; Xu, H. An energy-saving production of hydrogen peroxide via oxygen reduction for electro-Fenton using electrochemically modified polyacrylonitrile-based carbon fiber brush cathode. Sep. Purif. Technol. 2015, 156, 553–560. [Google Scholar] [CrossRef]
  139. Xia, G.; Lu, Y.; Xueli, G.; Gao, C.; Xu, H. Electro-Fenton Degradation of Methylene Blue Using Polyacrylonitrile-Based Carbon Fiber Brush Cathode. CLEAN Soil Ai Water 2014, 43, 229–236. [Google Scholar] [CrossRef]
  140. Gao, Y.; Zhu, W.; Li, Y.; Li, J.; Yun, S.; Huang, T. Novel porous carbon felt cathode modified by cyclic voltammetric electrodeposited polypyrrole and anthraquinone 2-sulfonate for an efficient electro-Fenton process. Int. J. Hydrogen Energy 2021, 46, 9707–9717. [Google Scholar] [CrossRef]
  141. Chen, S.; Tang, L.; Feng, H.; Zhou, Y.; Zeng, G.; Lu, Y.; Yu, J.; Ren, X.; Peng, B.; Liu, X. Carbon felt cathodes for electro-Fenton process to remove tetracycline via synergistic adsorption and degradation. Sci. Total Environ. 2019, 670, 921–931. [Google Scholar] [CrossRef]
  142. Wang, Y.; Liu, Y.; Wang, K.; Song, S.; Tsiakaras, P.; Liu, H. Preparation and characterization of a novel KOH activated graphite felt cathode for the electro-Fenton process. Appl. Catal. B Environ. 2015, 165, 360–368. [Google Scholar] [CrossRef]
  143. Want, Y.-T.; Tu, C.-H.; Lin, Y.-S. Application of Graphene and Carbon Nanotubes on Carbon Felt Electrodes for the Electro-Fenton System. Materials 2019, 12, 1698. [Google Scholar] [CrossRef] [Green Version]
  144. Özcan, A.; Şahin, Y.; Savaş Koparal, A.; Oturan, M.A. Carbon sponge as a new cathode material for the electro-Fenton process: Comparison with carbon felt cathode and application to degradation of synthetic dye basic blue 3 in aqueous medium. J. Electroanal. Chem. 2008, 616, 71–78. [Google Scholar] [CrossRef]
  145. Sopaj, F.; Oturan, N.; Pinson, J.; Podvorica, F.I.; Oturan, M.A. Effect of cathode material on electro-Fenton process efficiency for electrocatalytic mineralization of the antibiotic sulfamethazine. Chem. Eng. J. 2020, 384, 123249. [Google Scholar] [CrossRef]
  146. Ganiyu, S.O.; de Araújo, M.J.G.; de Araújo Costa, E.C.T.; Santos, J.E.L.; dos Santos, E.V.; Martínez-Huitle, C.A.; Pergher, S.B.C. Design of highly efficient porous carbon foam cathode for electro-Fenton degradation of antimicrobial sulfanilamide. Appl. Catal. B Environ. 2021, 283, 119652. [Google Scholar] [CrossRef]
  147. 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]
  148. Chu, Y.Y.; Qian, Y.; Wang, W.J.; Deng, X.L. A dual-cathode electro-Fenton oxidation coupled with anodic oxidation system used for 4-nitrophenol degradation. J. Hazard. Mater. 2012, 199–200, 179–185. [Google Scholar] [CrossRef]
  149. Chu, Y.; Su, H.; Lv, R.; Zhang, X. Enhanced electro-reduction of Fe3+ to Fe2+ by acidified carbon nanotube-modified graphite cathode and its application in a novel Fenton process for p-nitrophenol degradation. J. Water Process Eng. 2021, 40, 101912. [Google Scholar] [CrossRef]
  150. Zarei, M.; Beheshti Nahand, F.; Khataee, A.; Hasanzadeh, A. Removal of nalidixic acid from aqueous solutions using a cathode containing three-dimensional graphene. J. Water Process Eng. 2019, 32, 100978. [Google Scholar] [CrossRef]
  151. Zhou, W.; Meng, X.; Rajic, L.; Xue, Y.; Chen, S.; Ding, Y.; Kou, K.; Wang, Y.; Gao, J.; Qin, Y.; et al. “Floating” cathode for efficient H2O2 electrogeneration applied to degradation of ibuprofen as a model pollutant. Electrochem. Commun. 2018, 96, 37–41. [Google Scholar] [CrossRef]
  152. 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]
  153. Chu, Y.; Su, H.; Liu, C.; Zheng, X. Fabrication of sandwich-like super-hydrophobic cathode for the electro-Fenton degradation of cefepime: H2O2 electro-generation, degradation performance, pathway and biodegradability improvement. Chemosphere 2022, 286, 131669. [Google Scholar] [CrossRef]
  154. Karatas, O.; Gengec, N.A.; Gengec, E.; Khataee, A.; Kobya, M. High-performance carbon black electrode for oxygen reduction reaction and oxidation of atrazine by electro-Fenton process. Chemosphere 2021, 287, 132370. [Google Scholar] [CrossRef]
  155. Huang, A.; Zhi, D.; Zhou, Y. A novel modified Fe–Mn binary oxide graphite felt (FMBO-GF) cathode in a neutral electro-Fenton system for ciprofloxacin degradation. Environ. Pollut. 2021, 286, 117310. [Google Scholar] [CrossRef] [PubMed]
  156. Ma, B.; Lv, W.; Li, J.; Yang, C.; Tang, Q.; Wang, D. Promotion removal of aniline with electro-Fenton processes utilizing carbon nanotube 3D morphology modification of an Ag-loaded copper foam cathode. J. Water Process Eng. 2021, 43, 102295. [Google Scholar] [CrossRef]
  157. Dung, N.T.; Duong, L.T.; Hoa, N.T.; Thao, V.D.; Ngan, L.V.; Huy, N.N. A comprehensive study on the heterogeneous electro-Fenton degradation of tartrazine in water using CoFe2O4/carbon felt cathode. Chemosphere 2022, 287, 132141. [Google Scholar] [CrossRef] [PubMed]
  158. Paz, E.C.; Aveiro, L.R.; Pinheiro, V.S.; Souza, F.M.; Lima, V.B.; Silva, F.L.; Hammer, P.; Lanza, M.R.V.; Santos, M.C. Evaluation of H2O2 electrogeneration and decolorization of Orange II azo dye using tungsten oxide nanoparticle-modified carbon. Appl. Catal. B Environ. 2018, 232, 436–445. [Google Scholar] [CrossRef]
  159. Fdez-Sanromán, A.; Acevedo-García, V.; Pazos, M.; Sanromán, M.Á.; Rosales, E. Iron-doped cathodes for electro-Fenton implementation: Application for pymetrozine degradation. Electrochim. Acta 2020, 338, 135768. [Google Scholar] [CrossRef]
  160. 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]
  161. Yu, F.; Tao, L.; Cao, T. High yield of hydrogen peroxide on modified graphite felt electrode with nitrogen-doped porous carbon carbonized by zeolitic imidazolate framework-8 (ZIF-8) nanocrystals. Environ. Pollut. 2019, 255, 113119. [Google Scholar] [CrossRef]
  162. Cordeiro-Junior, P.J.; Kronka, M.S.; Goulart, L.A.; Veríssimo, N.C.; Mascaro, L.H.; dos Santos, M.C.; Bertazzoli, R.; de Vasconcelos Lanza, M.R. Catalysis of oxygen reduction reaction for H2O2 electrogeneration: The impact of different conductive carbon matrices and their physicochemical properties. J. Catal. 2020, 392, 56–68. [Google Scholar] [CrossRef]
  163. An, J.; Li, N.; Zhao, Q.; Qiao, Y.; Wang, S.; Liao, C.; Zhou, L.; Li, T.; Wang, X.; Feng, Y. Highly efficient electro-generation of H2O2 by adjusting liquid-gas-solid three phase interfaces of porous carbonaceous cathode during oxygen reduction reaction. Water Res. 2019, 164, 114933. [Google Scholar] [CrossRef] [PubMed]
  164. Ridruejo, C.; Alcaide, F.; Álvarez, G.; Brillas, E.; Sirés, I. On-site H2O2 electrogeneration at a CoS2-based air-diffusion cathode for the electrochemical degradation of organic pollutants. J. Electroanal. Chem. 2018, 808, 364–371. [Google Scholar] [CrossRef] [Green Version]
  165. Li, D.; Zheng, T.; Liu, Y.; Hou, D.; Yao, K.K.; Zhang, W.; Song, H.; He, H.; Shi, W.; Wang, L.; et al. A novel Electro-Fenton process characterized by aeration from inside a graphite felt electrode with enhanced electrogeneration of H2O2 and cycle of Fe3+/Fe2+. J. Hazard. Mater. 2020, 396, 122591. [Google Scholar] [CrossRef] [PubMed]
  166. Kubannek, F.; Turek, T.; Krewer, U. Modeling Oxygen Gas Diffusion Electrodesfor Various Technical Applications. Chem. Ing. Tech. 2019, 91, 720–733. [Google Scholar] [CrossRef] [Green Version]
  167. Wang, J.; Li, C.; Rauf, M.; Luo, H.; Sun, X.; Jiang, Y. Gas diffusion electrodes for H2O2 production and their applications for electrochemical degradation of organic pollutants in water: A review. Sci. Total Environ. 2021, 759, 143459. [Google Scholar] [CrossRef]
  168. Murawski, E.; Kananizadeh, N.; Lindsay, S.; Rao, A.M.; Popat, S.C. Decreased gas-diffusion electrode porosity due to increased electrocatalyst loading leads to diffusional limitations in cathodic H2O2 electrosynthesis. J. Power Sources 2021, 481, 228992. [Google Scholar] [CrossRef]
  169. Jiang, W.; Yang, H.C.; Yang, S.Y.; Horng, H.E.; Hung, J.C.; Chen, Y.C.; Hong, C.Y. Preparation and properties of superparamagnetic nanoparticles with narrow size distribution and biocompatible. J. Magn. Magn. Mater. 2004, 283, 210–214. [Google Scholar] [CrossRef]
  170. Lin, C.-C.; Lai, Y.-P.; Wu, K.-Y. A high-productivity process for mass-producing Fe3O4 nanoparticles by co-precipitation in a rotating packed bed. Powder Technol. 2021, 395, 369–376. [Google Scholar] [CrossRef]
  171. Lin, C.C.; Lee, C.Y. Adsorption of ciprofloxacin in water using Fe3O4 nanoparticles formed at low temperature and high reactant concentrations in a rotating packed bed with co-precipitation. Mater. Chem. Phys. 2020, 240, 122049. [Google Scholar] [CrossRef]
  172. Rao, M.N.; Sultana, R.; Kota, S.H. Hazardous Waste. Solid Hazard. Waste Manag. 2017, 1, 159–207. [Google Scholar] [CrossRef]
  173. Praveen, P.; Viruthagiri, G.; Mugundan, S.; Shanmugam, N. Structural, optical and morphological analyses of pristine titanium di-oxide nanoparticles—Synthesized via sol—Gel route. Spectrochim. Acta Part A Mol. Biomol. Spectrosc. 2014, 117, 622–629. [Google Scholar] [CrossRef]
  174. Dubey, R.S. Temperature-dependent phase transformation of TiO2 nanoparticles synthesized by sol-gel method. Mater. Lett. 2018, 215, 312–317. [Google Scholar] [CrossRef]
  175. Zanurin, A.; Johari, N.A.; Alias, J.; Ayu, H.M.; Redzuan, N.; Izman, S. Research progress of sol-gel ceramic coating: A review. Mater. Today Proc. 2021, 48, 1849–1854. [Google Scholar] [CrossRef]
  176. Demazeau, G. Solvothermal and hydrothermal processes: The main physico-chemical factors involved and new trends. Res. Chem. Intermed. 2010, 37, 107–123. [Google Scholar] [CrossRef]
  177. Krajian, H.; Abdallah, B.; Kakhia, M.; Alkafri, N. Hydrothermal growth method for the deposition of ZnO films: Structural, chemical and optical studies. Microelectron. Reliab. 2021, 125, 114352. [Google Scholar] [CrossRef]
  178. Dem, L.N.; Lyutin, V.I. Status of hydrothermal growth of bulk ZnO: Latest issues and advantages. J. Cryst. Growth 2008, 310, 993–999. [Google Scholar] [CrossRef]
  179. Yadav, S.; Sharma, A. Importance and challenges of hydrothermal technique for synthesis of transition metal oxides and composites as supercapacitor electrode materials. J. Energy Storage 2021, 44, 103295. [Google Scholar] [CrossRef]
  180. Guo, X.; Wang, K.; Xu, Y. Tartaric acid enhanced CuFe2O4-catalyzed heterogeneous photo-Fenton-like degradation of methylene blue. Mater. Sci. Eng. B 2019, 245, 75–84. [Google Scholar] [CrossRef]
  181. He, H.; Lu, J. Highly photocatalytic activities of magnetically separable reduced graphene oxide-CoFe2O4 hybrid nanostructures in dye photodegradation. Sep. Purif. Technol. 2017, 172, 374–381. [Google Scholar] [CrossRef]
  182. Boepple, M.; Zhu, Z.; Hu, X.; Weimar, U.; Barsan, N. Impact of heterostructures on hydrogen sulfidesensing: Example of core- shell CuO/CuFe2O4 nanostructures. Sens. Actuators B Chem. 2020, 321, 128523. [Google Scholar] [CrossRef]
  183. Liu, X.; Li, S.; Akinwolemiwa, B.; Hu, D.; Wu, T.; Peng, C. Low-crystalline transition metal oxide / hydroxide on MWCNT by Fenton-reaction-inspired green synthesis for lithium ion battery and OER electrocatalysis. Electrochim. Acta 2021, 387, 138559. [Google Scholar] [CrossRef]
  184. Cheng, Z.; Li, S.; Thang, T.; Gao, X.; Luo, S.; Guo, M. Biochar loaded on MnFe2O4 as Fenton catalyst for Rhodamine B removal: Characterizations, catalytic performance, process optimization and mechanism. Colloids Surf. A Physicochem. Eng. Asp. 2021, 631, 127651. [Google Scholar] [CrossRef]
  185. Zhou, T.; Huang, X.; Zhai, T.; Ma, K.; Zhang, H. Fabrication of novel three-dimensional Fe3O4 -based particles electrodes with enhanced electrocatalytic activity for Berberine removal. Chemosphere 2022, 287, 132397. [Google Scholar] [CrossRef] [PubMed]
  186. Yang, R.; Peng, Q.; Yu, B.; Shen, Y.; Cong, H. Yolk-shell Fe3O4@MOF-5 nanocomposites as a heterogeneous Fenton-like catalyst for organic dye removal. Sep. Purif. Technol. 2021, 267, 118620. [Google Scholar] [CrossRef]
  187. Zhu, X.; Zhang, L.; Zou, G.; Chen, Q.; Guo, Y.; Liang, S.; Hu, L.; North, M.; Xie, H. Carboxylcellulose hydrogel confined-Fe3O4 nanoparticles catalyst for Fenton-like degradation of Rhodamine B. Int. J. Biol. Macromol. 2021, 180, 792–803. [Google Scholar] [CrossRef]
  188. Jacobs, J.P.; Maltha, A.; Reinjtjes, J.; Drimal, J.; Ponec, V. Brongersm The surface of catalytically active spinels. Catalysis 1994, 147, 294–300. [Google Scholar] [CrossRef]
  189. So, H.L.; Lin, K.Y.; Chu, W. Triclosan removal by heterogeneous Fenton-like process: Studying the kinetics and surface chemistry of Fe3O4 as catalyst. J. Environ. Chem. Eng. 2019, 7, 103432. [Google Scholar] [CrossRef]
  190. Miller, A. Distribution of Cations in Spinels. Appl. Phys. 1959, 30, S24–S25. [Google Scholar] [CrossRef]
  191. Han, X.; Liu, S.; Huo, X.; Cheng, F.; Zhang, M.; Guo, M. Facile and large-scale fabrication of (Mg,Ni)(Fe,Al)2O4 heterogeneous photo-Fenton-like catalyst from saprolite laterite ore for effective removal of organic contaminants. J. Hazard. Mater. 2020, 392, 122295. [Google Scholar] [CrossRef]
  192. Wang, Y.; Tian, D.; Chu, W.; Li, M.; Lu, X. Nanoscaled magnetic CuFe2O4 as an activator of peroxymonosulfate for the degradation of antibiotics norfloxacin. Sep. Purif. Technol. 2019, 212, 536–544. [Google Scholar] [CrossRef]
  193. Amulya, M.A.S.; Nagaswarupa, H.P.; Kumar, M.R.A.; Ravikumar, C.R.; Kusuma, K.B.; Prashantha, S.C. Evaluation of bifunctional applications of CuFe2O4 nanoparticles synthesized by a sonochemical method. J. Phys. Chem. Solids 2021, 148, 109756. [Google Scholar] [CrossRef]
  194. Gao, J.; Wu, S.; Han, Y.; Tan, F.; Shi, Y.; Liu, M.; Li, X. 3D mesoporous CuFe2O4 as a catalyst for photo-Fenton removal of sulfonamide antibiotics at near neutral pH. J. Colloid Interface Sci. 2018, 524, 409–416. [Google Scholar] [CrossRef]
  195. Saravanakumar, B.; Ramachandran, S.P.; Ravi, G.; Ganesh, V.; Guduru, R.K.; Yuvakkumar, R. Electrochemical performances of monodispersed spherical CuFe2O4 nanoparticles for pseudocapacitive applications. Vacuum 2019, 168, 108798. [Google Scholar] [CrossRef]
  196. Wu, X.; Xia, F.; Nan, Z. Facile synthesis of double-mesoporous-shelled hollow spheres of Cu–CuFe2O4/SiO2 composite as excellent Fenton catalyst. Mater. Chem. Phys. 2020, 242, 122490. [Google Scholar] [CrossRef]
  197. Li, J.; Ren, Y.; Ji, F.; Lai, B. Heterogeneous catalytic oxidation for the degradation of p-nitrophenol in aqueous solution by persulfate activated with CuFe2O4 magnetic nano-particles. Chem. Eng. J. 2017, 324, 63–73. [Google Scholar] [CrossRef]
  198. Fontecha-Cámara, M.A.; Moreno-Castilla, C.; López-Ramón, M.V.; Álvarez, M.A. Mixed iron oxides as Fenton catalysts for gallic acid removal from aqueous solutions. Appl. Catal. B Environ. 2016, 196, 207–215. [Google Scholar] [CrossRef]
  199. Feng, Y.; Liao, C.; Shih, K. Copper-promoted circumneutral activation of H2O2 by magnetic CuFe2O4 spinel nanoparticles: Mechanism, stoichiometric efficiency, and pathway of degrading sulfanilamide. Chemosphere 2016, 154, 573–582. [Google Scholar] [CrossRef]
  200. Zhang, T.; Zhu, H.; Croue, J. Production of Sulfate Radical from Peroxymonosulfate Induced by a Magnetically Separable CuFe2O4 Spinel in Water: Efficiency, Stability, and Mechanism. Environ. Sci. Technol. 2013, 47, 2784–2791. [Google Scholar] [CrossRef]
  201. Suraj, P.; Kumar, V.; Thakur, C.; Ghosh, P. Taguchi optimization of COD removal by heterogeneous Fenton process using copper ferro spinel catalyst in a fixed bed reactor—RTD, kinetic and thermodynamic study. J. Environ. Chem. Eng. 2019, 7, 102859. [Google Scholar] [CrossRef]
  202. Ding, R.; Li, W.; He, C.; Wang, Y.; Liu, X.; Zhou, G.; Mu, Y. Oxygen vacancy on hollow sphere CuFe2O4 as an efficient Fenton-like catalysis for organic pollutant degradation over a wide pH range. Appl. Catal. B Environ. 2021, 291, 120069. [Google Scholar] [CrossRef]
  203. Yu, D.; Ni, H.; Wang, L.; Wu, M.; Yang, X. Nanoscale-con fined precursor of CuFe2O4 mediated by hyperbranched polyamide as an unusual heterogeneous Fenton catalyst for efficient dye degradation. J. Clean. Prod. 2018, 186, 146–154. [Google Scholar] [CrossRef]
  204. López-Ramón, M.V.; Álvarez, M.A.; Moreno-Castilla, C.; Fontecha-Cámara, M.A.; Yebra-Rodríguez, Á.; Bailón-García, E. Effect of calcination temperature of a copper ferrite synthesized by a sol-gel method on its structural characteristics and performance as Fenton catalyst to remove gallic acid from water. J. Colloid Interface Sci. 2018, 511, 193–202. [Google Scholar] [CrossRef]
  205. Xu, M.; Li, J.; Yan, Y.; Zhao, X.; Yan, J.; Zhang, Y.; Lai, B. Catalytic degradation of sulfamethoxazole through peroxymonosulfate activated with expanded graphite loaded CoFe2O4 particles. Chem. Eng. J. 2019, 369, 403–413. [Google Scholar] [CrossRef]
  206. Guo, X.; Li, H.; Zhao, S. Fast degradation of Acid Orange II by bicarbonate-activated hydrogen peroxide with a magnetic S-modified CoFe2O4 catalyst. Taiwan Inst. Chem. Eng. 2015, 55, 90–100. [Google Scholar] [CrossRef]
  207. Yao, Y.; Yang, Z.; Zhang, D.; Peng, W.; Sun, H.; Wang, S. Magnetic CoFe2O4-Graphene Hybrids: Facile Synthesis, Characterization, and Catalytic Properties. Ind. Eng. Chem. Res. 2012, 51, 6044–6051. [Google Scholar] [CrossRef]
  208. Hong, P.; Li, Y.; He, J.; Saeed, A.; Zhang, K. Rapid degradation of aqueous doxycycline by surface CoFe2O4/H2O2 system: Behaviors, mechanisms, pathways and DFT calculation. Appl. Surf. Sci. 2020, 526, 146557. [Google Scholar] [CrossRef]
  209. Feng, X.; Mao, G.Y.; Bu, F.X.; Cheng, X.L.; Jiang, D.M.; Jiang, J.S. Controlled synthesis of monodisperse CoFe2O4 nanoparticles by the phase transfer method and their catalytic activity on methylene blue discoloration with H2O2. Magn. Magn. Mater. 2013, 343, 126–132. [Google Scholar] [CrossRef]
  210. Singh, S.; Singhal, S. Transition metal doped cobalt ferrite nanoparticles: Efficient photocatalyst for photodegradation of textile dye. Mater. Today Proc. 2019, 14, 453–460. [Google Scholar] [CrossRef]
  211. Vinosha, P.A.; Das, S.J. Investigation on the role of pH for the structural, optical and magnetic properties of cobalt ferrite nanoparticles and its effect on the photo-fenton activity. Mater. Today Proc. 2018, 5, 8662–8671. [Google Scholar] [CrossRef]
  212. Kachi, W.; Al-Shammari, A.M.; Zainal, I.G. Cobalt Ferrite Nanoparticles: Preparation, characterization and salinized with 3-aminopropyl triethoxysilane. Energy Procedia 2019, 157, 1353–1365. [Google Scholar] [CrossRef]
  213. Sun, Z.; Liu, X.; Dong, X.; Zhang, X.; Tan, Y.; Yuan, F.; Zheng, S.; Li, C. Synergistic activation of peroxymonosulfate via in situ growth FeCo2O4 nanoparticles on natural rectorite: Role of transition metal ions and hydroxyl groups. Chemosphere 2021, 263, 127965. [Google Scholar] [CrossRef]
  214. Mohamed, S.G.; Tsai, Y.Q.; Chen, C.J.; Tsai, Y.T.; Hung, T.F.; Chang, W.S.; Liu, R.S. Ternary Spinel MCo2O4 (M = Mn, Fe, Ni, and Zn) Porous Nanorods as Bifunctional Cathode Materials for Lithium-O2 Batteries. Appl. Mater. Interfaces 2015, 7, 12038–12046. [Google Scholar] [CrossRef]
  215. Yadav, A.A.; Hunge, Y.M.; Kulkarni, S.B. Synthesis of multifunctional FeCo2O4 electrode using ultrasonic treatment for photocatalysis and energy storage applications. Ultrason. Sonochem. 2019, 58, 104663. [Google Scholar] [CrossRef] [PubMed]
  216. Zhang, T.; Ma, Q.; Zhou, M.; Li, C.; Sun, J.; Shi, W.; Ai, S. Degradation of methylene blue by a heterogeneous Fenton reaction catalyzed by FeCo2O4-N-C nanocomposites derived by ZIFs. Powder Technol. 2021, 383, 212–219. [Google Scholar] [CrossRef]
  217. Zhao, L.; Yang, D.; Ma, L.; Feng, X.; Ding, H. An efficient heterogeneous catalyst of FeCo2O4/g-C3N4 composite for catalytic peroxymonosulfate oxidation of organic pollutants under visible light. Colloids Surf. A 2021, 610, 125725. [Google Scholar] [CrossRef]
  218. Shan, Z.; Ma, F.; You, S.; Shan, L.; Kong, D.; Guo, H.; Cui, C. Enhanced visible light photo-Fenton catalysis by lanthanum-doping BiFeO3 for norfloxacin degradation. Environ. Res. 2023, 216, 114588. [Google Scholar] [CrossRef]
  219. Chen, X.; Qiao, J.; Wang, Z.; Sun, W.; Sun, K. Layered perovskites with exsolved Co-Fe nanoalloy as highly active and stable anodes for direct carbon solid oxide fuel cells. J. Alloys Compd. 2023, 940, 168872. [Google Scholar] [CrossRef]
  220. Torregrosa-Rivero, V.; Sánchez-Adsuar, M.-S.; Illán-Gómez, M.J. Exploring the effect of using carbon black in the sol-gel synthesis of BaMnO3 and BaMn0.7Cu0.3O3 perovskite catalysts for CO oxidation. Catal. Today, 2023; in press. [Google Scholar] [CrossRef]
  221. Nguyen, A.T.; Phung, V.D.; Mittova, V.O.; Ngo, H.D.; Vo, T.N.; Le Thi, M.L.; Nguyen, V.H.; Mittova, I.Y.; Le, M.L.P.; Ahn, Y.N.; et al. Fabricating nanostructured HoFeO3 perovskite for lithium-ion battery anodes via co-precipitation. Scr. Mater. 2022, 207, 114259. [Google Scholar] [CrossRef]
  222. Alrozi, R.; Zubir, N.A.; Bakar, N.F.A.; Ladewig, B.P.; Motuzas, J.; Bakar, N.H.H.A.; Wang, D.K.; da Costa, J.C.D. Functional role of B-site substitution on the reactivity of CaMFeO3 (M = Cu, Mo, Co) perovskite catalysts in heterogeneous Fenton-like degradation of organic pollutant. J. Taiwan Inst. Chem. Eng. 2023, 143, 104675. [Google Scholar] [CrossRef]
  223. Li, J.; Ma, W.; Zhong, D.; Li, K.; Ma, J.; Zhang, S.; Du, X. Oxygen vacancy concentration modulation of perovskite-based heterogeneous catalysts for Fenton-like oxidation of tetracycline. J. Clean. Prod. 2022, 362, 132469. [Google Scholar] [CrossRef]
  224. Xie, L.; Liu, X.; Chang, J.; Zhang, C.; Li, Y.; Zhang, H.; Zhan, S.; Hu, W. Enhanced redox activity and oxygen vacancies of perovskite triggered by copper incorporation for the improvement of electro-Fenton activity. Chem. Eng. J. 2022, 428, 131352. [Google Scholar] [CrossRef]
  225. Rusevova, K.; Köferstein, R.; Rosell, M.; Richnow, H.H.; Kopinke, F.D.; Georgi, A. LaFeO3 and BiFeO3 perovskites as nanocatalysts for contaminant degradation in heterogeneous Fenton-like reactions. Chem. Eng. J. 2014, 239, 322–331. [Google Scholar] [CrossRef]
  226. Zhao, H.; Cao, J.; Lv, H.; Wang, Y.; Zhao, G. 3D nano-scale perovskite-based composite as Fenton-like system for efficient oxidative degradation of ketoprofen. Catal. Commun. 2013, 41, 87–90. [Google Scholar] [CrossRef]
  227. Ho, J.H.; Li, Y.; Dai, Y.; Kim, T.; Wang, J.; Ren, J.; Yun, H.S.; Liu, X. Ionothermal synthesis of N-doped carbon supported CoMn2O4 nanoparticles as ORR catalyst in direct glucose alkaline fuel cell. Int. J. Hydrogen Energy 2021, 46, 20503–20515. [Google Scholar] [CrossRef]
  228. Estrada-Arriaga, E.B.; Guadarrama-Pérez, O.; Silva-Martínez, S.; Cuevas-Arteaga, C.; Guadarrama-Pérez, V.H. Oxygen reduction reaction (ORR) electrocatalysts in constructed wetland-microbial fuel cells: Effect of different carbon-based catalyst biocathode during bioelectricity production. Electrochim. Acta 2021, 370, 137745. [Google Scholar] [CrossRef]
  229. Gao, X.; He, L.; Yu, H.; Xie, F.; Yang, Y.; Shao, Z. The non-precious metal ORR catalysts for the anion exchange membrane fuel cells application: A numerical simulation and experimental study. Int. J. Hydrogen Energy 2020, 45, 23353–23367. [Google Scholar] [CrossRef]
  230. Liu, H.; Yang, D.H.; Wang, X.Y.; Zhang, J.; Han, B.H. N-doped graphitic carbon shell-encapsulated FeCo alloy derived from metal–polyphenol network and melamine sponge for oxygen reduction, oxygen evolution, and hydrogen evolution reactions in alkaline media. J. Colloid Interface Sci. 2021, 581, 362–373. [Google Scholar] [CrossRef] [PubMed]
  231. 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]
  232. Melchionna, M.; Fornasiero, P.; Prato, M. The Rise of Hydrogen Peroxide as the Main Product by Metal-Free Catalysis in Oxygen Reductions. Adv. Mater. 2019, 31, 1802920. [Google Scholar] [CrossRef]
  233. 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]
  234. Zhao, Q.; An, J.; Wang, X.; Li, N. In-situ hydrogen peroxide synthesis with environmental applications in bioelectrochemical systems: A state-of-the-art review. Int. J. Hydrogen Energy 2020, 46, 3204–3219. [Google Scholar] [CrossRef]
  235. Candia-Onfray, C.; Bollo, S.; Yáñez, C.; Escalona, N.; Marco, J.F.; Menéndez, N.; Salazar, R.; Recio, F.J. Nanostructured Fe-N-C pyrolyzed catalyst for the H2O2 electrochemical sensing. Electrochim. Acta 2021, 387, 138468. [Google Scholar] [CrossRef]
  236. Liu, Y.; Quan, X.; Fan, X.; Wang, H.; Chen, S. High-yield electrosynthesis of hydrogen peroxide from oxygen reduction by hierarchically porous carbon. Angew. Chem. Int. Ed. 2015, 54, 6837–6841. [Google Scholar] [CrossRef] [PubMed]
  237. Qiu, Z.; Ge, X.; Huang, N.; Zhou, S.; Zhang, J.; Xuan, J. Polyethylene oxide-polypropylene oxide -polyethylene oxide derived porous carbon materials with different molecular weights as ORR catalyst in alkaline electrolytes. Int. J. Hydrogen Energy 2021, 46, 2952–2959. [Google Scholar] [CrossRef]
  238. Inozemtseva, A.I.; Kataev, E.Y.; Frolov, A.S.; Amati, M.; Gregoratti, L.; Beranová, K.; Dieste, V.P.; Escudero, C.; Fedorov, A.; Tarasov, A.V.; et al. On the catalytic and degradative role of oxygen-containing groups on carbon electrode in non-aqueous ORR. Carbon N. Y. 2021, 176, 632–641. [Google Scholar] [CrossRef]
  239. Wang, X.; Sun, G.; Routh, P.; Kim, D.H.; Huang, W.; Chen, P. Heteroatom-doped graphene materials: Syntheses, properties and applications. Chem. Soc. Rev. 2014, 43, 7067–7098. [Google Scholar] [CrossRef] [Green Version]
  240. Eisenberg, D.; Stroek, W.; Geels, N.J.; Tanase, S.; Ferbinteanu, M.; Teat, S.J.; Mettraux, P.; Yan, N.; Rothenberg, G. A rational synthesis of hierarchically porous, N-doped carbon from Mg-based MOFs: Understanding the link between nitrogen content and oxygen reduction electrocatalysis. Phys. Chem. Chem. Phys. 2016, 18, 20778–20783. [Google Scholar] [CrossRef] [Green Version]
  241. Shibuya, R.; Kondo, T.; Nakamura, J. Active sites in nitrogen-doped carbon materials for oxygen reduction reaction. Carbon-Based Met. Catal. Des. Appl. 2018, 1–2, 227–249. [Google Scholar] [CrossRef]
  242. Zhou, W.; Rajic, L.; Meng, X.; Nazari, R.; Zhao, Y.; Wang, Y.; Gao, J.; Qin, Y.; Alshawabkeh, A.N. Efficient H2O2 electrogeneration at graphite felt modified via electrode polarity reversal: Utilization for organic pollutants degradation. Chem. Eng. J. 2019, 364, 428–439. [Google Scholar] [CrossRef]
  243. Zhu, J.; Xiao, X.; Zheng, K.; Li, F.; Ma, G.; Yao, H.; Wang, X.; Chen, Y. KOH-treated reduced graphene oxide: 100% selectivity for H2O2 electroproduction. Carbon 2019, 153, 6–11. [Google Scholar] [CrossRef]
  244. Wu, K.H.; Wang, D.; Lu, X.; Zhang, X.; Xie, Z.; Liu, Y.; Su, B.J.; Chen, J.M.; Su, D.S.; Qi, W.; et al. Highly Selective Hydrogen Peroxide Electrosynthesis on Carbon: In Situ Interface Engineering with Surfactants. Chem 2020, 6, 1443–1458. [Google Scholar] [CrossRef]
  245. Xu, H.; Guo, H.; Chai, C.; Li, N.; Lin, X.; Xu, W. Anodized graphite felt as an efficient cathode for in-situ hydrogen peroxide production and Electro-Fenton degradation of rhodamine B. Chemosphere 2022, 286, 131936. [Google Scholar] [CrossRef]
  246. Zahoor, A.; Christy, M.; Hwang, Y.J.; Lim, Y.R.; Kim, P.; Nahm, K.S. Improved electrocatalytic activity of carbon materials by nitrogen doping. Appl. Catal. B Environ. 2014, 147, 633–641. [Google Scholar] [CrossRef]
  247. Okamoto, Y. First-principles molecular dynamics simulation of O2 reduction on nitrogen-doped carbon. Appl. Surf. Sci. 2009, 256, 335–341. [Google Scholar] [CrossRef]
  248. Legarreta-Mendoza, A.; Flores-Holguín, N.; Lardizabal-Gutiérrez, D. A proposal based on quantum phenomena for the ORR mechanism on nitrogen-doped carbon-based electrocatalysts. Int. J. Hydrogen Energy 2019, 44, 12374–12380. [Google Scholar] [CrossRef]
  249. 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]
  250. Wang, Y.; Yi, M.; Wang, K.; Song, S. Enhanced electrocatalytic activity for H2O2 production by the oxygen reduction reaction: Rational control of the structure and composition of multi-walled carbon nanotubes. Chin. J. Catal. 2019, 40, 523–533. [Google Scholar] [CrossRef]
  251. Yu, Y.; You, S.; Du, J.; Zhang, P.; Dai, Y.; Liu, M.; Jiang, B.; Ren, N.; Zou, J. Ti3+-self-doped TiO2 with multiple crystal-phases anchored on acid-pickled ZIF-67-derived Co3O4@N-doped graphitized-carbon as a durable catalyst for oxygen reduction in alkaline and acid media. Chem. Eng. J. 2021, 403, 126441. [Google Scholar] [CrossRef]
  252. Li, Z.; Wang, W.; Zhou, M.; He, B.; Ren, W.; Chen, L.; Xu, W.; Hou, Z.; Chen, Y. In-situ self-templated preparation of porous core–shell Fe1−xS@N, S co-doped carbon architecture for highly efficient oxygen reduction reaction. J. Energy Chem. 2021, 54, 310–317. [Google Scholar] [CrossRef]
  253. Zhu, Y.; Deng, F.; Qiu, S.; Ma, F.; Zheng, Y.; Lian, R. Enhanced electro-Fenton degradation of sulfonamides using the N, S co-doped cathode: Mechanism for H2O2 formation and pollutants decay. J. Hazard. Mater. 2021, 403, 123950. [Google Scholar] [CrossRef]
  254. Soo, L.T.; Loh, K.S.; Mohamad, A.B.; Daud, W.R.W.; Wong, W.Y. An overview of the electrochemical performance of modified graphene used as an electrocatalyst and as a catalyst support in fuel cells. Appl. Catal. A Gen. 2015, 497, 198–210. [Google Scholar] [CrossRef]
  255. Zhang, L.; Xia, Z. Mechanisms of oxygen reduction reaction on nitrogen-doped graphene for fuel cells. J. Phys. Chem. C 2011, 115, 11170–11176. [Google Scholar] [CrossRef]
  256. Su, P.; Zhou, M.; Lu, X.; Yang, W.; Ren, G.; Cai, J. Electrochemical catalytic mechanism of N-doped graphene for enhanced H2O2 yield and in-situ degradation of organic pollutant. Appl. Catal. B Environ. 2019, 245, 583–595. [Google Scholar] [CrossRef]
  257. Su, P.; Zhou, M.; Song, G.; Du, X.; Lu, X. Efficient H2O2 generation and spontaneous [rad]OH conversion for in-situ phenol degradation on nitrogen-doped graphene: Pyrolysis temperature regulation and catalyst regeneration mechanism. J. Hazard. Mater. 2020, 397, 122681. [Google Scholar] [CrossRef] [PubMed]
  258. Qin, M.; Fan, S.; Wang, L.; Gan, G.; Wang, X.; Cheng, J.; Hao, Z.; Li, X. Oxygen and nitrogen co-doped ordered mesoporous carbon materials enhanced the electrochemical selectivity of O2 reduction to H2O2. J. Colloid Interface Sci. 2020, 562, 540–549. [Google Scholar] [CrossRef] [PubMed]
  259. Xie, Y.; Li, Y.; Huang, Z.; Zhang, J.; Jia, X.; Wang, X.S.; Ye, J. Two types of cooperative nitrogen vacancies in polymeric carbon nitride for efficient solar-driven H2O2 evolution. Appl. Catal. B Environ. 2020, 265, 118581. [Google Scholar] [CrossRef]
  260. Zhou, W.; Xie, L.; Gao, J.; Nazari, R.; Zhao, H.; Meng, X.; Sun, F.; Zhao, G.; Ma, J. Selective H2O2 electrosynthesis by O-doped and transition-metal-O-doped carbon cathodes via O2 electroreduction: A critical review. Chem. Eng. J. 2021, 410, 128368. [Google Scholar] [CrossRef]
  261. Iglesias, D.; Giuliani, A.; Melchionna, M.; Marchesan, S.; Criado, A.; Nasi, L.; Bevilacqua, M.; Tavagnacco, C.; Vizza, F.; Prato, M.; et al. N-Doped Graphitized Carbon Nanohorns as a Forefront Electrocatalyst in Highly Selective O2 Reduction to H2O2. Chem 2018, 4, 106–123. [Google Scholar] [CrossRef]
  262. Chen, J.Y.; Li, N.; Zhao, L. Three-dimensional electrode microbial fuel cell for hydrogen peroxide synthesis coupled to wastewater treatment. J. Power Sources 2014, 254, 316–322. [Google Scholar] [CrossRef]
  263. He, W.; Jiang, C.; Wang, J.; Lu, L. High-Rate Oxygen Electroreduction over Graphitic-N Species Exposed on 3D Hierarchically Porous Nitrogen-Doped Carbons. Angew. Chem. 2014, 53, 9503–9507. [Google Scholar] [CrossRef]
  264. Zhang, T.; Wu, J.; Wang, Z.; Wei, Z.; Liu, J.; Gong, X. Transfer of molecular oxygen and electrons improved by the regulation of C-N/C=O for highly efficient 2e-ORR. Chem. Eng. J. 2022, 433, 133591. [Google Scholar] [CrossRef]
  265. Chen, D.; Zhu, J.; Mu, X.; Cheng, R.; Li, W.; Liu, S.; Pu, Z.; Lin, C.; Mu, S. Nitrogen-Doped carbon coupled FeNi3 intermetallic compound as advanced bifunctional electrocatalyst for OER, ORR and zn-air batteries. Appl. Catal. B Environ. 2020, 268, 118729. [Google Scholar] [CrossRef]
  266. Gao, Y.; Zhu, W.; Wang, C.; Zhao, X.; Shu, M.; Zhang, J.; Bai, H. Enhancement of oxygen reduction on a newly fabricated cathode and its application in the electro-Fenton process. Electrochim. Acta 2020, 330, 135206. [Google Scholar] [CrossRef]
  267. Teng, Z.; Cai, W.; Sim, W.; Zhang, Q.; Wang, C.; Su, C.; Ohno, T. Applied Catalysis B: Environmental Photoexcited single metal atom catalysts for heterogeneous photocatalytic H2O2 production: Pragmatic guidelines for predicting charge separation. Appl. Catal. B Environ. 2021, 282, 119589. [Google Scholar] [CrossRef]
  268. Yeager, E. Dioxygen electrocatalysis: Mechanisms in relation to catalyst structure. J. Mol. Catal. 1986, 38, 5–25. [Google Scholar] [CrossRef]
  269. Chan, A.W.; Hoffmann, R.; Ho, W. Theoretical aspects of photoinitiated chemisorption, dissociation, and desorption of O2 on Pt(111). Langmuir 1992, 8, 1111–1119. [Google Scholar] [CrossRef]
  270. Viswanathan, V.; Hansen, H.A.; Rossmeisl, J.; Norskov, J. Unifying the 2e and 4e reduction of oxygen on metal surfaces. J. Phys. Chem. Lett. 2012, 3, 2948–2951. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  271. Kim, J.H.; Kim, Y.T.; Joo, S.H. Electrocatalyst design for promoting two-electron oxygen reduction reaction: Isolation of active site atoms. Curr. Opin. Electrochem. 2020, 21, 109–116. [Google Scholar] [CrossRef]
  272. Choi, C.H.; Kwon, H.C.; Yook, S.; Shin, H.; Kim, H.; Choi, M. Hydrogen peroxide synthesis via enhanced two-electron oxygen reduction pathway on carbon-coated pt surface. J. Phys. Chem. C 2014, 118, 30063–30070. [Google Scholar] [CrossRef]
  273. Perry, S.C.; Mavrikis, S.; Wang, L.; León, C.P. De Future perspectives for the advancement of electrochemical hydrogen peroxide production. Curr. Opin. Electrochem. 2021, 30, 100792. [Google Scholar] [CrossRef]
  274. Huang, Y.; Liu, W.; Kan, S.; Liu, P.; Hao, R.; Hu, H.; Zhang, J.; Liu, H.; Liu, M.; Liu, K. Tuning morphology and structure of Fe–N–C catalyst for ultra-high oxygen reduction reaction activity. Int. J. Hydrogen Energy 2020, 45, 6380–6390. [Google Scholar] [CrossRef]
  275. Daniel, G.; Kosmala, T.; Dalconi, M.C.; Nodari, L.; Badocco, D.; Pastore, P.; Lorenzetti, A.; Granozzi, G.; Durante, C. Upcycling of polyurethane into iron-nitrogen-carbon electrocatalysts active for oxygen reduction reaction. Electrochim. Acta 2020, 362, 137200. [Google Scholar] [CrossRef]
  276. Zhang, J.; Song, L.H.; Zhao, C.F.; Yin, X.P.; Zhao, Y.F. Co, N co-doped porous carbons as high-performance oxygen reduction electrocatalysts. Xinxing Tan Cailiao/N. Carbon Mater. 2021, 36, 209–218. [Google Scholar] [CrossRef]
  277. Chen, E.; Bevilacqua, M.; Tavagnacco, C.; Montini, T.; Yang, C.M.; Fornasiero, P. High surface area N/O co-doped carbon materials: Selective electrocatalysts for O2 reduction to H2O2. Catal. Today 2020, 356, 132–140. [Google Scholar] [CrossRef]
  278. Chen, J.; Wang, Z.; Mao, J.; Liu, C.; Chen, Y.; Lu, Z.; Feng, S.P. Bimetallic Ag–Cu nanosheets assembled flower-like structure for oxygen reduction reaction. J. Alloys Compd. 2021, 856, 157379. [Google Scholar] [CrossRef]
  279. Zhang, A.; Liu, Y.; Wu, J.; Zhu, J.; Cheng, S.; Wang, Y.; Hao, Y.; Zeng, S. Electrocatalytic selectivity to H2O2 enabled by two-electron pathway on Cu-deficient Au@Cu2-xS-CNTs electrocatalysts. Chem. Eng. J. 2023, 454, 140317. [Google Scholar] [CrossRef]
  280. Wang, B.; Jin, C.; Shao, S.; Yue, Y.; Zhang, Y.; Wang, S.; Chang, R.; Zhang, H.; Zhao, J.; Li, X. Electron-deficient Cu site catalyzed acetylene hydrochlorination. Green Energy Environ. 2022; 1–13, in press. [Google Scholar] [CrossRef]
  281. Noh, S.H.; Seo, M.H.; Ye, X.; Makinose, Y.; Okajima, T.; Matsushita, N.; Han, B.; Ohsaka, T. Design of an active and durable catalyst for oxygen reduction reactions using encapsulated Cu with N-doped carbon shells (Cu@N-C) activated by CO2 treatment. J. Mater. Chem. A 2015, 3, 22031–22034. [Google Scholar] [CrossRef]
  282. Lenarda, A.; Bevilacqua, M.; Tavagnacco, C.; Nasi, L.; Criado, A.; Vizza, F.; Melchionna, M.; Prato, M.; Fornasiero, P. Selective Electrocatalytic H2O2 Generation by Cobalt@N-Doped Graphitic Carbon Core–Shell Nanohybrids. ChemSusChem 2019, 12, 1664–1672. [Google Scholar] [CrossRef]
  283. Gao, J.; Yang, H.; Gao, J.; Yang, H.; Huang, X.; Hung, S.; Cai, W.; Jia, C. Enabling Direct H2O2 Production in Acidic Media through Rational Design of Transition Metal Single Atom Catalyst Enabling Direct H2O2 Production in Acidic Media through Rational Design of Transition Metal Single Atom Catalyst. Chem 2019, 6, 658–674. [Google Scholar] [CrossRef] [Green Version]
  284. Cao, K.W.; Huang, H.; Li, F.M.; Yao, H.C.; Bai, J.; Chen, P.; Jin, P.J.; Deng, Z.W.; Zeng, J.H.; Chen, Y. Co nanoparticles supported on three-dimensionally N-doped holey graphene aerogels for electrocatalytic oxygen reduction. J. Colloid Interface Sci. 2020, 559, 143–151. [Google Scholar] [CrossRef]
  285. Coliman, J.P.; Denisevich, P.; Konai, Y.; Marrocco, M.; Koval, C.; Anson, F.C. Electrode Catalysis of the Four-Electron Reduction of Oxygen to Water by Dicobalt Face-to-Face Porphyrins. J. Am. Chem. Soc. 1980, 102, 6027–6036. [Google Scholar] [CrossRef]
  286. Ferrara, M.; Bevilacqua, M.; Melchionna, M.; Criado, A.; Crosera, M.; Tavagnacco, C.; Vizza, F.; Fornasiero, P. Exploration of cobalt@N-doped carbon nanocomposites toward hydrogen peroxide (H2O2) electrosynthesis: A two level investigation through the RRDE analysis and a polymer-based electrolyzer implementation. Electrochim. Acta 2020, 364, 137287. [Google Scholar] [CrossRef]
  287. Cheng, X.; Dou, S.; Qin, G.; Wang, B.; Yan, P.; Taylor, T.; Yang, X. Rational design of highly selective nitrogen-doped Fe2O3-CNTs catalyst towards H2O2 generation in alkaline media. Int. J. Hydrogen Energy 2020, 45, 6128–6137. [Google Scholar] [CrossRef]
  288. Yue, D.; Qian, X.; Kan, M.; Fang, M.; Jia, J.; Yang, X.; Zhao, Y. A metal-free visible light active photo-electro-Fenton-like cell for organic pollutants degradation. Appl. Catal. B Environ. 2018, 229, 211–217. [Google Scholar] [CrossRef]
  289. Hartmann, M.; Kullmann, S.; Keller, H. Wastewater treatment with heterogeneous Fenton-type catalysts based on porous materials. J. Mater. Chem. 2010, 20, 9002–9017. [Google Scholar] [CrossRef]
  290. Li, Y.; Cao, W.X.; Zuo, X.J. O- and F-doped porous carbon bifunctional catalyst derived from polyvinylidene fluoride for sulfamerazine removal in the metal-free electro-Fenton process. Environ. Res. 2022, 212, 113508. [Google Scholar] [CrossRef]
  291. Yang, W.; Zhou, M.; Oturan, N.; Li, Y.; Su, P.; Oturan, M.A. Enhanced activation of hydrogen peroxide using nitrogen doped graphene for effective removal of herbicide 2,4-D from water by iron-free electrochemical advanced oxidation. Electrochim. Acta 2019, 297, 582–592. [Google Scholar] [CrossRef]
  292. Haider, M.R.; Jiang, W.L.; Han, J.L.; Sharif, H.M.A.; Ding, Y.C.; Cheng, H.Y.; Wang, A.J. In-situ electrode fabrication from polyaniline derived N-doped carbon nanofibers for metal-free electro-Fenton degradation of organic contaminants. Appl. Catal. B Environ. 2019, 256, 117774. [Google Scholar] [CrossRef]
  293. Qin, X.; Zhao, K.; Quan, X.; Cao, P.; Chen, S.; Yu, H. Highly efficient metal-free electro-Fenton degradation of organic contaminants on a bifunctional catalyst. J. Hazard. Mater. 2021, 416, 125859. [Google Scholar] [CrossRef]
  294. Chen, X.; Wang, L.; Sun, W.; Yang, Z.; Jin, J.; You, D.; Liu, G. Enhanced electrochemical advanced oxidation on boride activated carbon: The influences of boron groups. Electrochim. Acta 2021, 400, 139462. [Google Scholar] [CrossRef]
  295. Xie, F.; Gao, Y.; Zhang, J.; Bai, H.; Zhang, J.; Li, Z.; Zhu, W. A novel bifunctional cathode for the generation and activation of H2O2 in electro-Fenton: Characteristics and mechanism. Electrochim. Acta 2022, 430, 141099. [Google Scholar] [CrossRef]
  296. Zhang, Y.; Gao, M.; Wang, S.G.; Zhou, W.; Sang, Y.; Wang, X.H. Integrated electro-Fenton process enabled by a rotating Fe3O4/gas diffusion cathode for simultaneous generation and activation of H2O2. Electrochim. Acta 2017, 231, 694–704. [Google Scholar] [CrossRef]
  297. Cui, L.; Huang, H.; Ding, P.; Zhu, S.; Jing, W.; Gu, X. Cogeneration of H2O2 and [rad]OH via a novel Fe3O4/MWCNTs composite cathode in a dual-compartment electro-Fenton membrane reactor. Sep. Purif. Technol. 2020, 237, 116380. [Google Scholar] [CrossRef]
  298. Hu, J.; Wang, S.; Yu, J.; Nie, W.; Sun, J.; Wang, S. Duet Fe3C and FeNxSites for H2O2Generation and Activation toward Enhanced Electro-Fenton Performance in Wastewater Treatment. Environ. Sci. Technol. 2021, 55, 1260–1269. [Google Scholar] [CrossRef]
  299. Cao, P.; Quan, X.; Zhao, K.; Chen, S.; Yu, H.; Niu, J. Selective electrochemical H2O2 generation and activation on a bifunctional catalyst for heterogeneous electro-Fenton catalysis. J. Hazard. Mater. 2020, 382, 121102. [Google Scholar] [CrossRef]
  300. Ghasemi, M.; Khataee, A.; Gholami, P.; Soltani, R.D.C.; Hassani, A.; Orooji, Y. In-situ electro-generation and activation of hydrogen peroxide using a CuFeNLDH-CNTs modified graphite cathode for degradation of cefazolin. J. Environ. Manag. 2020, 267, 110629. [Google Scholar] [CrossRef]
  301. Cui, L.; Li, Z.; Li, Q.; Chen, M.; Jing, W.; Gu, X. Cu/CuFe2O4 integrated graphite felt as a stable bifunctional cathode for high-performance heterogeneous electro-Fenton oxidation. Chem. Eng. J. 2020, 127666. [Google Scholar] [CrossRef]
  302. Luo, T.; Feng, H.; Tang, L.; Lu, Y.; Tang, W.; Chen, S.; Yu, J.; Xie, Q.; Ouyang, X.; Chen, Z. Efficient degradation of tetracycline by heterogeneous electro-Fenton process using Cu-doped Fe@Fe2O3: Mechanism and degradation pathway. Chem. Eng. J. 2020, 382, 122970. [Google Scholar] [CrossRef]
  303. Sun, X.; Qi, H.; Sun, Z. Bifunctional nickel foam composite cathode co-modified with CoFe@NC and CNTs for electrocatalytic degradation of atrazine over wide pH range. Chemosphere 2022, 286, 131972. [Google Scholar] [CrossRef]
  304. Yu, D.; He, J.; Wang, Z.; Pang, H.; Li, L.; Zheng, Y.; Chen, Y.; Zhang, J. Mineralization of norfloxacin in a CoFe–LDH/CF cathode-based heterogeneous electro-fenton system: Preparation parameter optimization of the cathode and conversion mechanisms of H2O2 to ·OH. Chem. Eng. J. 2021, 417, 129240. [Google Scholar] [CrossRef]
  305. Li, Y.; Wang, C.; Pan, S.; Zhao, X.; Liu, N. Mn doping improves in-situ H2O2 generation and activation in electro-Fenton process by Fe/Mn@CC cathode using high-temperature shock technique. Chemosphere 2022, 307, 136074. [Google Scholar] [CrossRef] [PubMed]
  306. Xiao, F.; Wang, Z.; Fan, J.; Majima, T.; Zhao, H.; Zhao, G. Selective Electrocatalytic Reduction of Oxygen to Hydroxyl Radicals via 3-Electron Pathway with FeCo Alloy Encapsulated Carbon Aerogel for Fast and Complete Removing Pollutants. Angew. Chem. Int. Ed. 2021, 60, 10375–10383. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Scheme of the Fenton process. In general, the multiple reactions involved can be summarized by Equations (1) and (2) [40,41].
Figure 1. Scheme of the Fenton process. In general, the multiple reactions involved can be summarized by Equations (1) and (2) [40,41].
Catalysts 13 00674 g001
Figure 2. Schematic of the OH generation by MnFe2O4@C-NH2. Adapted from [56].
Figure 2. Schematic of the OH generation by MnFe2O4@C-NH2. Adapted from [56].
Catalysts 13 00674 g002
Figure 3. (a) SEM images of Chem-Sch, Bio-Sch, Fe3O4/Sch, Fe3O4/Sch/C, (b) TEM of Fe3O4/Sch and Fe3O4/Sch/C, (c) HRTEM of Fe3O4/Sch/C, and (d) elemental mapping of different elements C, Fe, S, and O recorded from Fe3O4/Sch/C [60].
Figure 3. (a) SEM images of Chem-Sch, Bio-Sch, Fe3O4/Sch, Fe3O4/Sch/C, (b) TEM of Fe3O4/Sch and Fe3O4/Sch/C, (c) HRTEM of Fe3O4/Sch/C, and (d) elemental mapping of different elements C, Fe, S, and O recorded from Fe3O4/Sch/C [60].
Catalysts 13 00674 g003
Figure 4. Effect of the addition of carbon in the promoted reduction of Fe3+ to Fe2 [104].
Figure 4. Effect of the addition of carbon in the promoted reduction of Fe3+ to Fe2 [104].
Catalysts 13 00674 g004
Figure 5. SEM images of carbon foam with low (a) and high magnification (b) [146].
Figure 5. SEM images of carbon foam with low (a) and high magnification (b) [146].
Catalysts 13 00674 g005
Figure 6. SEM images of “G-Smooth” cathode (a), “G-Rough” cathode (b), “G/CNT 10%” cathode (c), “G/CNT 20%” cathode (d), and “G/CNT 50%” cathode (e); Water contact angle images of the cathodes (f) [153].
Figure 6. SEM images of “G-Smooth” cathode (a), “G-Rough” cathode (b), “G/CNT 10%” cathode (c), “G/CNT 20%” cathode (d), and “G/CNT 50%” cathode (e); Water contact angle images of the cathodes (f) [153].
Catalysts 13 00674 g006
Figure 7. SEM images of (a) the unmodified graphite felt (raw GF) cathode, (b) the unmodified graphite felt (FMBO–GF) cathode, and (c) TOC removal efficiency and the mineralization current efficiency (MCE) of raw GF and FMBO–GF in 2 h under 25 mg/L initial pollutant concentration, pH 7, and 2 mA/cm2 current density [155].
Figure 7. SEM images of (a) the unmodified graphite felt (raw GF) cathode, (b) the unmodified graphite felt (FMBO–GF) cathode, and (c) TOC removal efficiency and the mineralization current efficiency (MCE) of raw GF and FMBO–GF in 2 h under 25 mg/L initial pollutant concentration, pH 7, and 2 mA/cm2 current density [155].
Catalysts 13 00674 g007
Figure 8. Schematic of a gas diffusion electrode adapted from [166].
Figure 8. Schematic of a gas diffusion electrode adapted from [166].
Catalysts 13 00674 g008
Figure 9. SEM (a) and TEM (b) of CuFe2O4 hollow spheres [202].
Figure 9. SEM (a) and TEM (b) of CuFe2O4 hollow spheres [202].
Catalysts 13 00674 g009
Figure 10. Schematic illustration of the effect of N-contained surface groups in the generation and activation of H2O2. Adapted from [256].
Figure 10. Schematic illustration of the effect of N-contained surface groups in the generation and activation of H2O2. Adapted from [256].
Catalysts 13 00674 g010
Figure 11. Synthesis of N-doped graphitized carbon nanohorns and TEM images at each stage of the synthesis [261].
Figure 11. Synthesis of N-doped graphitized carbon nanohorns and TEM images at each stage of the synthesis [261].
Catalysts 13 00674 g011
Figure 12. ORR pathways depend on the geometric structure of catalysts. Adapted from [271].
Figure 12. ORR pathways depend on the geometric structure of catalysts. Adapted from [271].
Catalysts 13 00674 g012
Figure 13. Effect of a carbon coating on Pt/C catalysts surface by acetylene CVD, and H2O2 selectivity depending on the CVD operation time for Pt/C catalysts. Adapted from [272].
Figure 13. Effect of a carbon coating on Pt/C catalysts surface by acetylene CVD, and H2O2 selectivity depending on the CVD operation time for Pt/C catalysts. Adapted from [272].
Catalysts 13 00674 g013
Figure 14. (a) Possible heterogeneous EF catalytic mechanisms using Fe3O4/MWCNTs/CB GDE dual−composite. (b) TEM and (c) SEM images of Fe3O4/MWCNTs/CB composite [297].
Figure 14. (a) Possible heterogeneous EF catalytic mechanisms using Fe3O4/MWCNTs/CB GDE dual−composite. (b) TEM and (c) SEM images of Fe3O4/MWCNTs/CB composite [297].
Catalysts 13 00674 g014
Table 2. Strategies and improvements in the development of electro-Fenton Catalysts.
Table 2. Strategies and improvements in the development of electro-Fenton Catalysts.
CatalystStrategyImprovementRef.
Fe (NO3)3·9H2OAddition of MWCNT and complexation on the surface of the MWCNT through the surface groups.
  • The reaction rate was 23 times higher than in traditional Fenton.
  • Faster Fe3+/Fe2+ cycle (Demonstrated by the generation of OH and HO2, combined with an accelerated decomposition of the contaminant).
  • Better catalytic activity by MWCNT carboxylates groups. Accelerating the intermediate step from Fe-OOH to Fe2+.
[102]
FeCl3·6H2OAddition of carbon materials (powdered activated carbon and carbon nanotubes).
  • The reduction from Fe3+ to Fe2+ was accelerated. Due to the reducing power of carbon materials (especially carbonyl/quinone groups).
  • Carbon nanotubes had a lower rate of decomposition of H2O2 than powdered activated carbon.
[103]
Fe (NO3)3·9H2OAddition of hydrothermally prepared carbon.
  • Promoted the Fe3+/Fe2+ cycle through the transfer of electrons from carbonaceous material to Fe3+.
  • Hydroxyl groups on the surface of hydrothermal carbon strongly affect the reduction of Fe3+.
[104]
Fe2O3@FeBFeB coating to Fe2O3 core.
  • An efficient Fe3+/Fe2+ cycle was achieved, thanks to the oxidizing and constantly ceding electrons to Fe3+.
  • The Fe2+ could also generate H2O2 in situ by the activation pathway of a single electron of O2.
[115]
FeCl3·6H2OAddition of granular activated carbon.
  • The formation of Fe2+ was accelerated, which improved the catalytic oxidation of the target contaminant.
[105]
FerrihydriteAddition of biocarbon and S. oneidensis.
  • Faster Fe3+ reduction rate.
[106]
Montmorillonite (Fe/Cu-MMT)Preparation of Fe–Cu bimetallic active sites intercalated on montmorillonite.
  • The bleaching efficiency of the contaminant was higher, thanks to the electrons involved in the redox reaction of both metals, which promoted the rate of mutual oxidation.
  • A decrease in the leaching of pure metal ions of Fe and Cu, by presenting good stability with the support.
[111]
Fe–Mn/SAPO-18Fe–Mn bimetallic active sites supported on SAPO-18.
  • Improved degradation efficiency of the contaminant, compared to the single-metal catalysts (Fe and Mn) supported on the zeolite.
[112]
Fe3O4–SepSepiolite support for Fe3O4 nanoparticles.
  • Improved adsorption of the catalyst and, therefore, the in situ degradation of the contaminant on the surface of Fe3O4-Sep.
[113]
γ-Fe2O3–SiO2Mesoporosity creation and Fe3+ doping of maghemite/silica microspheres.
  • The substitution of microporosity or mesoporosity led to a greater discoloration of the effluent to be treated.
[114]
QuFe@ACFActive carbon fiber support for ferric hydroxyquinoline.
  • Free electrons in the carbon fibers were able to drive electron transfer to ferric 8-hydroxyquinoline to promote the Fe3+/Fe2+ cycle faster.
[107]
Fe-CFe-doped carbon xerogel.
  • Improved dispersion of Fe0 species.
[116]
CF/rGOReduced graphene oxide support for copper ferrite.
  • The redox efficiency improved due to the two metal ions (Cu2+/Cu+ and Fe3+/Fe2+).
  • Synergistic effect between the nanotube structure of the copper ferrite together with the mesoporosity of the support.
[108]
Fe–N/CCarbon xerogel support for heteroatoms (Fe, N).
  • Improved mass transfer to catalytic active sites.
  • Increased activation of H2O2 due to the coexistence of Fe and N.
[109]
CoFe2O4/NOMXerogel support from the natural organic matrix (NOM) for CoFe2O4.
  • Increased surface area, increasing exposure to catalytically active sites.
  • Improved electron transfer between metal ion pairs, which increased the speed of the redox cycle (Co2+/Co3+ and Fe2+/Fe3+).
[110]
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

Fajardo-Puerto, E.; Elmouwahidi, A.; Bailón-García, E.; Pérez-Cadenas, A.F.; Carrasco-Marín, F. From Fenton and ORR 2e-Type Catalysts to Bifunctional Electrodes for Environmental Remediation Using the Electro-Fenton Process. Catalysts 2023, 13, 674. https://doi.org/10.3390/catal13040674

AMA Style

Fajardo-Puerto E, Elmouwahidi A, Bailón-García E, Pérez-Cadenas AF, Carrasco-Marín F. From Fenton and ORR 2e-Type Catalysts to Bifunctional Electrodes for Environmental Remediation Using the Electro-Fenton Process. Catalysts. 2023; 13(4):674. https://doi.org/10.3390/catal13040674

Chicago/Turabian Style

Fajardo-Puerto, Edgar, Abdelhakim Elmouwahidi, Esther Bailón-García, Agustín Francisco Pérez-Cadenas, and Francisco Carrasco-Marín. 2023. "From Fenton and ORR 2e-Type Catalysts to Bifunctional Electrodes for Environmental Remediation Using the Electro-Fenton Process" Catalysts 13, no. 4: 674. https://doi.org/10.3390/catal13040674

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