Next Article in Journal
A Prolyl Endopeptidase from Flammulina velutipes Degrades Celiac Disease-Inducing Peptides in Grain Flour Samples
Next Article in Special Issue
A Critical Review of Photo-Based Advanced Oxidation Processes to Pharmaceutical Degradation
Previous Article in Journal
The 3D-Printing Fabrication of Multichannel Silicone Microreactors for Catalytic Applications
Previous Article in Special Issue
Photocatalytic Concrete Developed by Short Seedless Hydrothermal Method for Water Purification
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Recent Advances in the Development of Novel Iron–Copper Bimetallic Photo Fenton Catalysts

by
Gabriela N. Bosio
1,
Fernando S. García Einschlag
1,
Luciano Carlos
2,* and
Daniel O. Mártire
1,*
1
Instituto de Investigaciones Fisicoquímicas Teóricas y Aplicadas (INIFTA), Departamento de Química, Facultad de Ciencias Exactas, Universidad Nacional de la Plata, Consejo Nacional de Investigaciones Científicas y Técnicas (CONICET), La Plata 1900, Argentina
2
Instituto de Investigación y Desarrollo en Ingeniería de Procesos, Biotecnología y Energías Alternativas, PROBIEN (CONICET-UNCo), Universidad Nacional del Comahue, Neuquén 8300, Argentina
*
Authors to whom correspondence should be addressed.
Catalysts 2023, 13(1), 159; https://doi.org/10.3390/catal13010159
Submission received: 23 November 2022 / Revised: 21 December 2022 / Accepted: 4 January 2023 / Published: 10 January 2023

Abstract

:
Advanced oxidation processes (AOPs) have been postulated as viable, innovative, and efficient technologies for the removal of pollutants from water bodies. Among AOPs, photo-Fenton processes have been shown to be effective for the degradation of various types of organic compounds in industrial wastewater. Monometallic iron catalysts are limited in practical applications due to their low catalytic activity, poor stability, and recyclability. On the other hand, the development of catalysts based on copper oxides has become a current research topic due to their advantages such as strong light absorption, high mobility of charge carriers, low environmental toxicity, long-term stability, and low production cost. For these reasons, great efforts have been made to improve the practical applications of heterogeneous catalysts, and the bimetallic iron–copper materials have become a focus of research. In this context, this review focuses on the compilation of the most relevant studies on the recent progress in the application of bimetallic iron–copper materials in heterogeneous photo–Fenton-like reactions for the degradation of pollutants in wastewater. Special attention is paid to the removal efficiencies obtained and the reaction mechanisms involved in the photo–Fenton treatments with the different catalysts.

Graphical Abstract

1. Introduction

In recent years, the rapid development of the industrial sector has resulted in numerous effluents that contain toxic substances and affect the environment in a dangerous way [1]. Some of these pollutants, due to their chemical nature, are resistant to the conventional physical and biological processes commonly used in wastewater treatment plants [2]. For this reason, advanced oxidation processes (AOPs) have been postulated as viable, innovative, and efficient technologies for the removal of these compounds from water bodies [3]. Among the AOPs, the heterogeneous Fenton and photo–Fenton processes have been shown to be effective for the degradation of various types of organic compounds in industrial wastewater [4,5].
Equations (1)–(5) are the main pathways involved in the Fenton processes. Reaction 1 is a key step because it generates the strongly oxidizing radical HO, while regeneration of Fe2+ is the rate–determining step under dark conditions and in the absence of alternative Fe3+–reducing species (Equation (2)) [6,7].
F e 2 + + H 2 O 2 F e 3 + +   H O + H O ( k = 40 80   M 1 s 1 )
F e 3 + + H 2 O 2 F e 2 + + H O 2   +   H + ( k = 9.1 × 10 7   M 1 s 1 )
F e 2 + + H O F e 3 + +   H O ( k = 2.5 5 × 10 8   M 1 s 1 )
F e 3 + + H O 2       F e 2 + +   H + +   O 2 ( k = 0.33 2.1 × 10 6   M 1 s 1 )
F e 2 + +   H O 2 F e 3 + + H O 2 ( k = 0.72 1.5 × 10 6   M 1 s 1 )
Equations (1)–(5) are the most important steps in the dark Fenton chemistry because H2O2 is consumed and Fe2+ is regenerated from Fe3+ by these reactions. The HO and HO2 radicals are the main species involved in the oxidation of pollutants. In addition to Fe(II), other transition metal ions can also promote similar processes, which are then referred to as Fenton–like or Fenton–type. For example, both oxidation states of copper can react with H2O2 to form HO2 and HO radicals (Equations (6) and (7)) [8].
C u 2 + + H 2 O 2 C u + + H O 2     +   H + ( k = 1.15 × 10 6   M 1 s 1 )
C u + + H 2 O 2 C u 2 + +   H O + H O ( k = 1.0 × 10 4   M 1 s 1 )
Under suitable conditions, copper can act as a better catalyst compared to iron, as it is able to form temporary complexes with oxidation products and rapidly convert Cu+ to Cu2+ and vice versa [9,10,11]. Since the oxidation products do not form stable complexes with copper, reactive coordination sites remain available for continuous catalytic cycling. Therefore, copper not only provides a better redox cycle but is also active in the mineralization of organic matter. However, a major disadvantage of copper–based catalysts is the high excess of H2O2 required to maintain catalytic activity, which increases treatment costs [12]. This disadvantage has often been mitigated by the use of bimetallic composites of copper and iron [9,13,14,15].
Since Reaction 1 proceeds much faster than Reaction 2, the Fe2+ ions are consumed faster than generated, so that a large amount of ferric hydroxide is formed as sludge during the process, which causes additional problems in separation and disposal [6,16,17]. Fenton (H2O2/Fe2+) and Fenton–like (H2O2/Fe3+) processes can be significantly enhanced under ultraviolet (UV)/visible irradiation (λ < 600 nm). The UV region of the electromagnetic spectrum extends from about 100 nm to 400 nm, while the visible region extends from about 400 nm to 760 nm [18]. Photoirradiation prevents the accumulation of Fe3+ ions in the system since Fe2+ ions are regenerated from Fe3+ by photoreduction [17,19,20,21]. The improvement in pollutant removal rates achieved with the photo–Fenton process compared to the Fenton process can be explained by the following contributions [6,22,23,24]: (i) the generation of HO and Fe2+ via the photolysis of iron (III) hydroxo complexes (λ < 580 nm) (Equations (8) and (9)) [25,26], followed by the participation of Fe2+ in Equation (1) to generate further HO radicals [27], (ii) the photolysis of H2O2 when λ < 310 nm is used, and (iii) the photolysis of Fe(III) chelates (Fe3L) formed between Fe3+ and the organic substrate, its degradation intermediates or other ligands present in the reaction medium (Equation (10)) [6]. However, the photo–Fenton process still has some disadvantages, such as the narrow pH range (2.8–3.5), poor switching between Fe2+ and Fe3+, and sludge formation. Therefore, many efforts have been made to develop catalysts that can operate efficiently at circumneutral pH conditions [28,29,30,31].
[ F e ( O H ) ] 2 + + h ν F e 2 + + H O
[ F e ( H 2 O ) ] 3 + + h ν F e 2 + + H O + H +
[ F e 3 + L ] + h ν [ F e 3 + L ] + L
In the heterogeneous Fenton process, solid iron catalysts (such as the iron minerals magnetite (Fe3O4) or hematite (α–Fe2O3)) are used instead of aqueous Fe2+, to catalyze the generation of hydroxyl radicals in acidic or near–neutral media [32,33]. The mechanism of this process starts with the adsorption of H2O2 on the surface of Fe–based catalysts [34,35]. Then, the formation of a surface complex (≡Fe2+–H2O2 and/or ≡Fe3+–H2O2) precursor occurs (Equations (11) and (12)). Subsequently, the Fe2+–H2O2 complex generates ≡Fe3+ and the surface–bound HO radical by intramolecular electron transfer (Reaction 13) [36]. The ≡Fe3+–H2O2 complex can be converted to ≡Fe2+ and OOH (Equation (14)) [36], and the HOO can further regenerate ≡Fe2+ (Equation (15)) [37]. The HO can either attack the adsorbed organic compounds in the vicinity or oxidize the non–adsorbed organic compounds.
F e 2 + + H 2 O 2     F e 2 + H 2 O 2
F e 3 + + H 2 O 2     F e 3 + H 2 O 2
F e 2 + H 2 O 2   F e 3 + +   H O + H O
F e 3 + H 2 O 2     F e 2 + +   H O O + H +
F e 3 + + H O O     F e 2 + +   O 2 + H +
Monometallic iron catalysts (ZVI, Fe2O3, Fe3O4, and Fe(OOH)) are limited in practical applications due to their low catalytic activity, poor stability, and recyclability [38]. On the other hand, copper–based oxides have become the priorities of research into novel photocatalysts, due to their advantages such as strong light absorption, high carrier mobility, non–toxicity, environmental friendliness, long–term stability, and low production cost. In particular, copper has the property of a Lewis acid, which can create a localized acidic microenvironment by interacting with iron species. The effect of expanding the pH range of the system is achieved by the localized acidic microenvironment when bimetallic Fe–Cu photocatalysts are used [39]. In this context, great efforts have been made to improve the practical applications of heterogeneous catalysts, and bimetallic oxides (composite oxides of iron and another metallic element) are an active research area. Similarly, bimetallic catalysts composed of iron and copper are advantageous for the reduction of Fe3+ in Fenton–like reactions. Previous reports have suggested that the cooperation between the redox pairs of iron (Fe3+/Fe2+) and copper (Cu2+/Cu+) [40,41] can accelerate electron transfer at the interface to promote the rapid reduction of Fe3+. Han et al. [42] explained the synergistic effect due to the reduced ΔE for the Fe3+/Fe2+ redox cycle obtained for the bimetallic Fe–Cu catalyst compared to the monometallic Fe catalyst. Sun et al. [43] suggested that the Fe2+ species of the Fe–Cu bimetallic catalyst is mainly regenerated by the reaction between Fe3+ and Cu+ (Equation (16)) rather than by the reduction of Fe3+ by H2O2 (Equation (2)). Thus, the coexistence of Fe and Cu on the surface of a catalyst can accelerate the process of electron transfer in the reaction environment and create a suitable condition for the generation of reactive radical species.
C u + + F e 3 +   C u 2 + +   F e 2 + Δ E 0 = 0.6   V
In summary, for the above reasons, bimetallic iron–copper materials are likely to perform better as heterogeneous Fenton and photo–Fenton catalysts compared to monometallic iron or copper materials. In this context, this review focuses on the compilation of the most relevant studies on the recent advances in the application of bimetallic iron– copper materials as Fe and Cu sources in heterogeneous photo–Fenton–like reactions for the degradation of pollutants in wastewater. Although the differences in the model pollutants and the experimental conditions make an evaluation difficult, we used as central parameters for the comparisons between the different materials, on the one hand, the combinations “achieved efficiency/treatment time” (which can be found in the tables) and, on the other hand, the dominant mechanisms in each study. For clarity, the results in the next sections are classified into the following groups according to the chemical composition of the catalysts (Figure 1): (i) CuFe2O4, (ii) mixed ferrites containing Fe and Cu, (iii) CuFeO2 and CuFeS2 materials, (iv) Fe–Cu oxide composites, (v) Metal–Organic Frameworks (MOF) based on Fe and Cu.

2. CuFe2O4

Ferrites are a broad class of iron–bearing oxides that include spinel ferrites, perovskites, ilmenite (FeTiO3), and hexagonal ferrites (or hexaferrites). In general, spinel ferrites can be described as ferrites with the formula MFe2O4, where M is a divalent metal cation (such as Fe, Co, Zn, Ni, Cu, or others) [44]. Due to their thermal stability [45], unique structural [46], optical [47], magnetic [48], electrical, and dielectric properties [49], spinel ferrites have broad potential technological applications in photoluminescence [50], photocatalysis [51], biosensor development [52], magnetic drug delivery [53], corrosion protection [54], antimicrobial agents [55], and biomedicine (hyperthermia) [56]. In particular, CuFe2O4 has an inverse spinel structure with 8 Cu2+ ions in octahedral sites and 16 Fe3+ ions evenly distributed between the tetrahedral (A) and octahedral (B) sites [57].
The Tauc’s equation was used to determine the optical band gap (Eg) of copper ferrite [58]:
α h ν = B ( h ν E g ) m
where α is the energy–dependent absorption coefficient, is the photon energy, B is an energy independent constant, and the factor m depends on the nature of the electron transition and is equal to 2 or 1/2 for the direct and indirect transition band gaps, respectively. Assuming a direct semiconductor, Eg can be obtained from the extrapolation of the linear section of Tauc’s plot of (αhν)0.5 vs. to the energy axis. In comparison with other spinel ferrites such as CoFe2O4 (Eg = 2.7 eV) and NiFe2O4 NPs (Eg = 2.2 eV) [59], CuFe2O4 NPs have a relatively small band gap of (1.7–1.9 eV) which allows them to use freely available energy in the form of sunlight to degrade pollutants in water [59]. Since CuFe2O4 is a narrow band gap material, it can be successfully used in type II or Z–scheme heterojunction photocatalysts [60].
Wei et al. [39] prepared magnetic Fe–Cu materials by the sol–gel method. Solutions containing 0–40% Cu2+ were prepared separately to obtain catalysts with different iron and copper ratios. The molar ratio of iron and copper had significant effects on the surface and structural properties of the materials. These materials were used for the removal of pyridine under UV irradiation in a Fenton–like system. The effects of various parameters, such as the molar ratio of Fe and Cu, catalyst and H2O2 dosage, initial pH, and initial pyridine concentration were studied. The best degradation performance was obtained for the M–40% material prepared from the 40% Cu2+ solution, which consisted of Fe2O3 and CuFe2O4. This co–doped Fe–Cu catalyst showed high activity, good stability, and easy recovery for the degradation of pyridine when the catalyst loading was 0.90 g L−1, the H2O2 dosage was 832.50 mg L−1, the initial pH was 7, and the initial pyridine concentration was 100.00 mg L−1. The degradation efficiency of pyridine was more than 99% within 30 min and the TOC removal efficiency was 97.4% within 50 min. Therefore, the M–40% catalyst could be a potential material for wastewater treatment, which could extend the active pH range to facilitate recycling and reuse. The proposed reaction mechanism involves the reduction of Cu2+ to Cu+ at the surface of the photocatalyst (Equation (6)), and the reaction of Cu+ with H2O2 to form hydroxyl radicals and Cu2+ (Equation (7)). In addition, HO is obtained from Cu2+ and water under UV irradiation (Equation (17)). Part of Cu2+ is reduced to Cu+ by O2 according to Equation (18). Furthermore, the monovalent copper ions can also be oxidized by H2O2 to produce Cu3+, which reacts with water molecules under UV light to form hydroxyl radicals (Equations (19) and (20)). Direct evidence for the occurrence of Equation (19) was obtained by EPR spectroscopy using the spin–trapping approach [61]. It is very likely that this reaction occurs at the surface of the photocatalyst. The complete mechanism is shown in Figure 2.
C u 2 + + H 2 O +   h ν C u + + H O + H +
C u 2 + + O 2   C u + +   O 2
C u + + H 2 O 2   C u 3 + +   2   H O
C u 3 + + H 2 O   +   h ν     C u 2 + + H O + H +
Silva et al. [62] prepared mixed iron and copper oxides by the modified Pechini´s method. The samples were calcined at different temperatures and used for the removal of Methylene Blue dye (MB) by solar photo–Fenton catalysis. The results of X–ray diffraction analysis showed that hematite and copper ferrite phases were mainly responsible for the high efficiency of the photochemical reaction. MB removal from the aqueous solution was carried out in the presence of H2O2 (300 mg L−1), with a 1.0 g L−1 catalyst, at a neutral pH, and under solar irradiation. One of the catalysts was more efficient than pure iron (Fe2O3) and copper oxides (CuO), indicating the synergistic effect produced by combining the two metals on the same material. The authors proposed a mechanism for the regeneration of the active sites that satisfactorily explained the experimental results.
Cao et al. [63] prepared CuFe2O4 composites via combustion solution synthesis. The authors studied the effect of Cu content on the synthesized composites. The composite with 18 wt.% Cu had the best photo–Fenton activity. This nanocomposite, which was composed of CuFe2O4 and CuO, could degrade 40 mg L−1 MB, 20 mg L−1 Rhodamine B (RhB), and 20 mg L−1 methyl orange (MO) in 40, 30, and 50 min, respectively. In addition, the composite showed superparamagnetic behavior and could be recycled.
Guo et al. [40] synthesized self–assembled hollow nanospheres of CuFe2O4 using the solvothermal method. The catalytic activity of the nanospheres was evaluated by the degradation of MB. In addition, these authors compared the properties of CuFe2O4 particles prepared by different methods. It was found that the good performance of the solid catalyst depends on both the large specific surface area and the degree of the optical response. The increase in photoelectric response and conductivity are beneficial for the improvement of catalytic performance. Transient photo–current experiments showed that the samples obtained by the solvothermal method exhibited a better optical response in the on–off cycle under light conditions.
Leichtweis et al. [64] prepared a novel composite catalyst by doping CuFe2O4 nanoparticles in the malt bagasse biochar. Malt bagasse (the malted barley residue) is the main residue obtained in the manufacture of beer. Composites with different ratios of malt biochar and CuFe2O4 were produced. The composites had lower band gap energy than CuFe2O4, which increased the photocatalytic activity. At 60 min of heterogeneous photo–Fenton treatment (pH 3), visible light tests showed that pure CuFe2O4 removed only 39% of the RhB color, while the composites removed up to 88% of the dye. A total of 100% color removal (pH 3) was achieved at 10- and 20-min reaction times for 10 and 50 mg L−1 of dye under solar irradiation. Tests performed in the presence of specific radical scavengers showed that HO, O2, and h+ were the predominant reactive species involved in the degradation of the dye.
Jiang et al. [65] prepared a magnetic Bi2WO6/CuFe2O4 catalyst to be used in the removal of the antibiotic tetracycline hydrochloride (TCH). The obtained catalyst achieved 92.1% TCH (20 mg L−1) degradation efficiency in a photo–Fenton–like system and a mineralization performance of 50.7% and 35.1% for TCH and a raw secondary effluent from a wastewater treatment plant, respectively. The excellent performance was attributed to the fact that photogenerated electrons accelerated the conversion Fe(III)/Fe(II) and Cu(II)/Cu(I), which increased the reaction rates of Fe(II)/Cu(I) with H2O2 and generated abundant HO radicals for pollutant oxidation. EPR assays confirmed that O2 and HO were mainly responsible for THC degradation in dark and photo–Fenton–like systems, respectively.
The low efficiency of photo–Fenton processes at a neutral pH is mainly due to the precipitation of iron and can therefore be prevented by the proper addition of iron complexing agents, such as tartaric acid [66]. Guo et al. [67] prepared CuFe2O4 particles by the sol–gel method and investigated the effects of tartaric acid (TA) on the degradation of MB in the presence of CuFe2O4 and H2O2 under light irradiation. The results showed that the introduction of TA increased the decolorization rate of MB decolorization from 52.0% to 92.1% within 80 min. The contribution of O2 was only 10%, whereas that of HO was about 88%. The enhancement of MB degradation in the presence of TA was explained by the complexation of Fe3+ and Cu2+ with TA (Equation (21)), followed by a photo–induced ligand to metal charge transfer process (Equation (22)). In the literature, similar equations with iron species and other organic radicals have been proposed to explain the photo–Fenton mechanism [66]. The photo–induced intramolecular charge transfer of the copper (II) complex of tartaric acid has already been described in the mechanism of the catalytic reduction of Cr(VI) by tartaric acid under the irradiation of simulated sunlight [68]. The excited species Fe2+/Cu+–(TA) further produced Fe2+/Cu+ and TA radicals according to Equation (23). Moreover, the presence of molecular oxygen also accelerated the transformation of the chemical state of the metal ions (see Equation (24)).
F e 3 + / C u 2 + + T A   F e 3 + ( T A ) / C u 2 + ( T A )
F e 3 + ( T A ) / C u 2 + ( T A ) + h ν     F e 2 + ( T A ) / C u + ( T A )
F e 2 + ( T A ) / C u + ( T A )     F e 2 + / C u +   +   ( T A )
F e 2 + ( T A ) / C u + ( T A )     +       O 2   F e 3 + ( T A ) / C u + ( T A )     +   O 2
Rocha et al. [69] prepared CuFe2O4 from copper recycled from spent lithium–ion batteries. The photocatalytic properties were analyzed by monitoring the decolorization of MB in a heterogeneous photo–Fenton process in the presence of solar radiation. The decolorization efficiency was 96.1% in 45 min of reaction Formic and acetic acids were detected as degradation products of MB by ion chromatography.
Lin and Lu [70] developed CuFe2O4 nanoparticles decorated on partially reduced graphene oxides (CuFe2O4@rGO), to selectively and efficiently cleave lignin model compounds into value–added aromatic chemicals via a sunlight–assisted heterogeneous Fenton process. Controlled oxidative cleavage of the lignin model compound enabled the production of high–value aromatic chemicals, guaiacol and 2–methoxy–4–propylphenol, in high yields, instead of complete mineralization of the lignin model compound. The partially reduced graphene oxides serve as the large size support to accommodate and immobilize the CuFe2O4 nanoparticles for easy recycling of the catalyst, to attract the lignin model compounds through π–π stacking and hydrogen bonding for efficient cleavage reaction, and to accelerate the transport of photo–induced electrons for better charge separation and thus higher photocatalytic activities.
Membrane separation offers many advantages in wastewater treatment, including high efficiency, low energy consumption, and ease of operation [71]. However, membrane fouling always lowers the separation efficiency and shortens the membrane life, which greatly hinders the application of membrane technology. For this reason, Wang et al. [72] combined photo–Fenton and membrane processes for water treatment (Figure 3). These authors synthesized CuFe2O4 particles and doped them into the PVDF@CuFe2O4 membranes. The photo–Fenton process can degrade various foulants through the generation of hydroxyl radicals, which improves the filtration performance of the membranes.
The PVDF@CuFe2O4 membrane (1.0% CuFe2O4) showed significant improvement in both permeability and separation, with a tripling of flux and doubling of rejection compared to the values obtained by the membrane filtration process itself. In addition, the PVDF@CuFe2O4 nanofiltration membrane showed excellent stability and reusability after repeated tests over fifteen cycles.
Recently reported studies on the use of CuFe2O4 as photo–Fenton photocatalysts are summarized in Tables below.

3. Mixed Ferrites Containing Fe and Cu

As already mentioned, among the iron–based materials used as heterogeneous photo–Fenton catalysts, ferrites have attracted much attention because they are chemically and thermally stable magnetic materials [73,74]. Most importantly, ferrites have a narrow band gap, which enables them to efficiently utilize the visible region of solar energy in photocatalysis [75]. MnFe2O4 has a large specific surface area, good biocompatibility, and excellent magnetic properties [76,77]. However, mixed ferrites have better catalytic behavior compared to single ferrites [78]. In particular, doping with Cu2+ could improve the optical properties of MnFe2O4, which is due to the lattice defects in the spinel structure generated as a consequence of the smaller ionic radius of Cu2+. In addition, doping with other metals can also introduce oxygen vacancies, leading to a significant improvement in catalytic performance due to structural distortions [79]. Meena et al. [80] prepared Ce–doped MnFe2O4 by a low–temperature solution combustion synthesis using Oxalyl Dihydrazine (ODH) as fuel and reported that the band gap decreased after Ce–doping compared to that of pure MnFe2O4. Moreover, copper, manganese, and iron have been shown to have synergistic effects [81]. Thus, Cu2+–doped MnFe2O4 could further improve the catalytic activity because the different radii of metal ions in the spinel structure may lead to some defects and distortions, which result in the production of oxygen vacancies that are beneficial for the formation of reactive oxygen species (ROS) [82].
Yang et al. [83] synthesized porous Cu0.5Mn0.5Fe2O4 nanoparticles by a co–precipitation process and a subsequent high–temperature annealing treatment method. Cu0.5Mn0.5Fe2O4 nanoparticles exhibited much higher catalytic activity towards the degradation of bisphenol A (BPA) by the activation of H2O2 under UV light irradiation compared with CuFe2O4 and MnFe2O4 nanoparticles. The authors proved, through radical scavenger and EPR/DMPO experiments, that the hydroxyl radical is involved and plays a critical role in the presence of H2O2. They also proposed a possible BPA degradation pathway.
Sun et al. [82] employed Cu0.8Mn0.2Fe2O4 in the presence of H2O2 and achieved a great improvement in the degradation efficiency of TCH, which was due to the fact that H2O2 was activated by the Fe, Mn, and Cu ions on the Cu0.8Mn0.2Fe2O4 surface to generate HO radicals. Two possible mechanisms for H2O2 activation by Cu0.8Mn0.2Fe2O4 at pH = 3 and pH = 11 were proposed. Firstly, Cu0.8Mn0.2Fe2O4 could be excited by visible light to generate electrons (e) and holes (h+), then H2O2 could react with the generated e to produce HO (Equation (25)). Photoelectrons migrating to the surface of the catalyst immediately react with the metal ions of the catalyst, which not only promotes the circulation of Fe, Mn, and Cu components but also strengthens the synergistic effect between them (Equations (16), (26) and (27)). When H2O2 came into contact with Fe, Mn, and Cu ions, the Fenton reaction was triggered immediately and TCH was mineralized by these reactive species into small molecule compounds.
H 2 O 2 +     e     H O + H O
M n 3 + + F e 2 +     M n 2 + +   F e 3 +
M n 3 + + C u +       M n 2 + +   C u 2 +
Wang et al. [84] used a p–n heterostructured nano–photocatalyst based on the p–type Cu2ZnSnS4 (CZTS) nanosheets, which were successfully assembled on the surface of ZnFe2O4 (ZFO) nanospheres, forming CZTS/ZFO p–n heterostructures. These p–n heterostructures could not only efficiently expand the spectral response and promote photo–induced charge separation, but also increase the specific surface areas for photocatalytic and photo–Fenton reactions. All these factors resulted in the p–n hetero–structured CZTS/ZFO nano–photocatalyst with significantly enhanced photocatalytic activity for the degradation of MO in the presence of H2O2 with visible light irradiation, compared to pure ZFO. This behavior was due to the synergistic enhancement effects of the CZTS/ZFO p–n heterostructure combined with the photo–Fenton mechanism. (Figure 4).
Shi et al. [85] fabricated a multi–functional honeycomb ceramic plate by coating a layer of CuFeMnO4 on the surface of a cordierite (magnesium iron aluminum cyclosilicate) material. The honeycomb structure was beneficial for light trapping and energy recycling and thus improved the solar–to–water evaporation efficiency. The CuFeMnO4 coating layer acted as both the photothermal material for the solar–driven water evaporation process and the catalyst for the removal of volatile organic compounds (VOCs) via the heterogeneous photo–Fenton reaction. With the integration of the photo–Fenton reaction into the solar distillation process, clean distilled water was produced with efficient removal of the potential VOCs from the contaminated water sources.
All the results described in this section are summarized in Table 1.

4. CuFeO2 and CuFeS2 Materials

Bimetallic Cu(II) and Fe(II) oxides and sulfides capable of activating H2O2 for the degradation of organic pollutants have become a focus of research. Among these catalysts, CuFeO2 has gained wide concern due to its simple synthesis, low cost, interesting catalytic activity, and high chemical stability. In addition, due to its narrow band gap and cooperative impact of Fe3+/Fe2+ and Cu2+/Cu+ redox cycles, CuFeO2 materials usually exhibit significant visible–light absorption and a good performance for H2O2 activation.
Schmachtenberg et al. [86] obtained delafossite–type powders by conventional hydrothermal (CuFeO2) and microwave–assisted hydrothermal (CuFeO2–MW) routes. Both materials were tested as potential catalysts in the photo–Fenton reaction under visible light. An experimental design was used for optimizing the degradation efficiency of the dye Reactive Red 141 and assessing the effects of the operating variables pH, catalyst loading, and oxidant concentration. Both a substantial reduction in the treatment times and a significant efficiency improvement were observed for the catalyst prepared by microwave irradiation (i.e., about 98% of dye degradation at 30 min) in comparison with the material obtained by the conventional method (i.e., 84% at 150 min). The latter difference was ascribed to the fact that CuFeO2–MW samples presented higher values of surface area and pore volume, as well as smaller band–gap energy in comparison with conventional CuFeO2. In addition, under optimal conditions (i.e., pH = 3.0, [CuFeO2–MW] = 0.25 g L−1, and [H2O2] = 8 mmol L−1) 80% of TOC removal was achieved at 180 min. Finally, reusability tests conducted with the CuFeO2–MW catalyst showed only a marginal loss of decolorization efficiency after four cycles (from 98% to 93%) and total concentrations of leached Fe and Cu ions of less than 1.0 wt.% of the catalyst content.
Liu et al. [87] prepared stoichiometric CuFeO2 microcrystals with high crystallinity and single crystal phase via the optimized hydrothermal process. The authors presented a detailed characterization including the analysis of crystal structure, optical properties, and photoelectrochemical behavior. TCH was used as a model compound to evaluate the activity of CuFeO2 under different conditions. Since the microcrystals showed a nonnegligible photocatalytic activity in the absence of H2O2, the authors proposed that the photo–generated electrons in the CB of CuFeO2 could interact with dissolved oxygen and the photo–generated holes in the valence band react with H2O, thus forming active radicals such as O2 and HO, respectively. On the other hand, in the presence of H2O2, the material showed some catalytic activity for the Fenton reaction under dark conditions, which was enhanced more than four times upon illumination. The heterogeneous photo–Fenton–like treatment of TCH, carried out at pH 8 and using a Xe–Lamp in a quartz cell, showed conversion degrees of about 80 and 90% within the first 60 and 120 min, respectively. Photo–Fenton assays performed in the presence of different radical scavengers indicated that HO is not directly responsible for the degradation, but that the photogenerated holes and O2 are the predominant reactive species.
Da Silveira Salla et al. [88] evaluated the catalytic properties of CuFeS2–MW through BPA degradation tests conducted at a near–neutral pH and using simulated visible light. Results showed a remarkable enhancement of the catalytic efficiency (i.e., about 90% in the first 15 min) for the catalyst obtained in the presence of citrate, with a rate about 10 times faster than that of CuFeS2 prepared without citrate. This behavior suggested that the presence of citric acid accelerates the photo–conversion of Fe3+ to Fe2+, thus increasing the generation of HO and the efficiency of the overall process. Reusability assays showed that, after the fourth cycle, the degradation rate remained higher than 95% and the TOC decrease was 76.6% after 60 min of reaction. In addition, concentrations of Fe and Cu leached into the solution were lower than 0.5% of the total amounts of Fe and Cu present in the catalyst at the beginning of the reaction. To assess the main reactive species, tests were conducted in the presence of different ROS scavengers (i.e., t–butanol for HO and p–benzoquinone for O2). Results showed that BPA degradation may be ascribed to the generation of HO. As expected, due to the low selectivity of the HO radicals, the degradation efficiency was affected by the presence of other chemical species in the real effluent such as carbonates/bicarbonates, nitrates/nitrites, sulfates, phosphates, and chlorides. Using LC–MS analysis, the authors identified several hydroxylated derivates resulting from the initial reaction stages. These primary intermediates are likely to undergo ring–opening reactions to form typical aliphatic acids as degradation by–products, which may further undergo a series of oxidation steps that finally lead to complete mineralization.
More recently, Da Silveira Salla et al. [89] prepared a chalcopyrite powder by a microwave–assisted method (CuFeS2–MW), which exhibited higher catalytic activity than that obtained with the material prepared by the conventional method (CuFeS2). Both materials were synthesized in the presence of citric acid, which is capable of enhancing the photoreduction of Fe(III). Tartrazine dye was used as a model compound for the degradation by the photo–Fenton reaction under visible irradiation at pH 3.0. CuFeS2–MW reached 99.1% of tartrazine decolorization at 40 min and 87.3% of mineralization at 150 min, the decolorization rate being twice as fast as that obtained with CuFeS2. The latter difference was attributed to the synergy between higher crystallinity and increased amount of Fe2+ on the CuFeS2–MW surface when compared to CuFeS2. Reusability tests performed on both CuFeS2–MW and CuFeS2 showed that, after five cycles, the decolorization efficiencies remained at 93.6 and 69.6%, respectively. Moreover, the highest concentrations of leached iron were less than 1% of the total amount of iron present in the catalysts. The addition of t–butanol decreased 20–fold the tartrazine decolorization rate, while p–benzoquinone did not inhibit tartrazine decolorization. These results show that HO radicals are the key species for tartrazine removal and that O2 radicals do not play a significant role in the degradation mechanism.
Cai et al. [90] anchored CuFeO2 on a Manganese residue (MR) through mechanical activation (MA) for obtaining Fe–Cu@SiO2/starch–derived carbon composites. The material was tested as a heterogeneous catalyst for the photo–Fenton treatment of TC using H2O2 and visible light. Under optimal conditions (i.e., 15 mM of H2O2 and pH 7.0) 100% of TC conversion (50 mg L−1) was achieved within 40 min. Moreover, the observed decay constants were 4.00, 2.77, and 2.14 times higher than those of Cu/SC, MAMR–Fe3O4@SiO2/SC, and MR–Fe–Cu@SiO2/SC, respectively. Reutilization tests showed good material stability since the photocatalytic performance of MAMR–Fe–Cu@SiO2/SC changed from 99.2% to 96.3% after five cycles and metal leaching was below 0.1 mg L−1 for Cu, Fe, and Mn. Based on XPS analyses before and after material usage, the authors proposed that the efficiency of the prepared catalyst is closely related to the interaction of Cu2+/Cu+, Fe3+/Fe2+, and Mn3+/Mn2+ couples. The study of reactive species through the use of scavengers and ESR showed that both HO and O2 contribute to TC degradation in the MAMR–Fe–Cu@SiO2/SC + H2O2 + visible light system, while surface photogeneration of electrons and holes seems to play a negligible role.
Xin et al. [38] synthesized CuFeO2/biochar composites for heterogeneous photo–Fenton–like processes (HPF–like) via the hydrothermal method. Biochar prevented agglomeration and avoided the usage of added reductants. Moreover, the introduction of biochar had several advantages, such as enhancing visible–light absorption, narrowing the bandgap of CuFeO2, and partially suppressing the recombination between photoelectron and hole pairs. By applying the design of experiments and the surface response methodology the authors studied the effects of the operating conditions on TC degradation efficiency. The results showed that the optimum parameters were 220 mg L−1 of catalysts, 22 mM of H2O2, and pH 6.4, with a pH dependence relatively gentle in the range of 4.0 to 8.0. Under optimal conditions, the efficiency of TC degradation in HPF–like systems was 96.7% after 120 min of treatment. Furthermore, the catalyst exhibited excellent stability since the activity only decreased by 3.9% after five utilization cycles and the total dissolved concentrations of both iron and copper ions at 120 min were below 0.02 mg L−1. Active species scavenging experiments were carried out by using p–benzoquinone, silver nitrate, methyl alcohol, and ammonium oxalate as the scavengers of O2, photoelectrons, HO, and photogenerated holes, respectively. Interestingly, ammonium oxalate did not decrease the catalytic performance but accelerated the TC degradation rate. The latter effect was attributed to a lower photoelectron quenching rate, which could promote Fe3+/Fe2+ and Cu2+/Cu+ cycles, thus significantly improving H2O2 activation. In addition, ESR experiments allowed the identification of HO as the predominant active species, whereas photoelectrons and O2, were auxiliary species for TC degradation. The mechanistic findings are schematically summarized in Figure 5.
It is worth mentioning that the authors also explored the TC degradation intermediate products by HPLC–MS and proposed plausible transformation pathways. On the other hand, Xin et al. [91] also studied the heterogeneous visible–light Photo–electro–Fenton (H–VL–PEF) treatment of TC using an undivided photoelectrochemical cell in the presence of CuFeO2/biochar particles. A nitrogen/oxygen self–doped biomass porous carbon cathode was used as a gas diffusion electrode (GDE) to efficiently produce H2O2, while a Xe–Lamp (filter > 420 nm) was used as a visible irradiation source for CuFeO2/biochar particles. The performances of different treatments (including electro–catalysis, photo–catalysis, photo–electro–catalysis, electro–Fenton, and H–VL–PEF) were evaluated for comparable experimental setups. The H–VL–PEF showed excellent performance, achieving the highest TC degradation rate and the best TOC removal with the lowest energy consumption. The authors analyzed the effect of several operational parameters and found that the full decomposition of TC (20 to 200 mg L−1) was attained within 60–70 min under optimal conditions (i.e., 100 mg L−1 CuFeO2/biochar, 80 mA cm−2 of current density, and pH 5.0). Reutilization tests showed negligible metal leaching and small efficiency losses after 10 cycles. Experiments performed in the presence of selective scavengers of active species as well as ESR measurements indicated that HO radicals are the main species responsible for TC degradation, whereas O2 radicals play a subsidiary role. The schematic representation of the involved processes is presented in Figure 6. In addition, based on the intermediates detected by HPLC–MS analyses, the authors proposed general pathways for TC transformation and used a software tool for predicting, through QSAR techniques, the evolution of sample toxicity with the treatment time.
More recently, Xin et al. [92] achieved a further enhancement in the performance of photo–electro–Fenton (PEF) degradation of TC by introducing modifications to the latter treatment strategy. Three GDE composites were fabricated by mixing different weight ratios of a nitrogen/oxygen self–doped porous biochar (NO/PBC) cathode for H2O2 electrogeneration with a NO/PBC–supported CuFeO2 (CuFeO2–NO/PBC) catalyst for H2O2 activation. The NO/PBC was prepared by the pyrolysis method, whereas CuFeO2–NO/PBC was prepared by the hydrothermal method without additional chemical reductants. The PEF treatment of TC was carried out in an undivided quartz reactor with a Pt anode and using visible light (Xe–Lamp, cutoff > 420 nm). The authors analyzed the effects of the NO/PBC to CuFeO2–NO/PBC ratio, the current density, and initial pH on system performance, which depends on the compromise between H2O2 formation, parallel reactions such as 4–electron oxygen reduction reaction (ORR) and H2 evolution, and H2O2 decomposition. Using a NO/PBC to CuFeO2–NO/PBC ratio of 1:1, a current density of 80 mA cm−2, and pH 5, an efficiency of 96.1% was achieved for TC (20 mg L−1) degradation at 180 min in the PEF treatment. The latter value is twice the rate observed in EF and an order of magnitude higher than that recorded under anodic oxidation. The degradation efficiency at 180 min of PEF treatment steadily declined from 96.1% at pH 5.0 to 46.8% at pH 11.0. After five cycles of material use, the PEF treatment efficiency was higher than 80%, the amounts of leached Fe and Cu were negligible, and surface hydrophobicity remained practically constant. Assays in the presence of radical scavengers showed that HO and O2 played the primary and auxiliary roles for tetracycline degradation, respectively. The authors proposed that despite the photoinduced hole could not effectively participate in TC degradation, the photogenerated electrons reacted with H2O2, O2, and Fe3+/Cu2+ to form HO, O2, and Fe2+/Cu+, respectively, thus promoting an efficient TC transformation. Finally, from the reaction intermediates detected by HPLC–MS, QSAR tools, and experiments of E. coli growth inhibition, a general picture of TC oxidation pathways and toxicity evolution was outlined.
All the results described in this section are summarized in Table 2.

5. Fe–Cu Oxide Composites

In this section, we present the results obtained with other types of heterogeneous catalysts consisting of mixtures of iron and copper oxides and mixed oxides of defined composition. Heterojunctions formed by different copper and iron oxides and/or sulfides decrease the recombination rate of photogenerated e and h+ [93,94].
Mansoori et al. [95] prepared an iron–copper oxide impregnated NaOH–activated biochar (Fe–Cu/ABC) through the pyrolysis of activated biochar, followed by the impregnation method. The catalytic activity of the bimetallic catalyst was tested for ciprofloxacin (CIP) degradation through a heterogeneous photo–electro–Fenton process) at a natural pH. The heterogeneous catalyst exhibited remarkable catalytic activity and showed great stability and structural integrity for five cycles. Furthermore, from a practical point of view, the catalyst exhibited an acceptable performance by oxidizing CIP dissolved in various water matrices such as tap water, river water, and a real sample of wastewater. The intermediates by–products of CIP were determined, and a plausible degradation mechanism was proposed. The adsorption of as–generated H2O2 onto the well–developed surface of the mesoporous bimetallic catalyst facilitated the reduction of Fe3+ to Fe2+ and Cu2+ to Cu+. The co–existence of Fe and Cu on the surface of activated biochar accelerates the process of electron transfer in the reaction and enhances the production of reactive species.
Iron oxide NPs supported on the surface of hydroxylated diamond exhibited photocatalytic activity and stability for the visible light–assisted Fenton reaction even working at pH values around 6 and demonstrated superiority as supports compared to other alternative solids such as activated carbon, graphite, carbon nanotubes, and even the benchmark semiconductor TiO2 [96]. In particular, Manickam–Periyaraman, et al. [97] prepared bimetallic Fe20Cu80 NPs supported on various surface hydroxylated diamond NPs (D) and compared them to analogous catalysts based on graphite. The assays were conducted at pH 6 for the heterogeneous Fenton degradation of phenol (Ph) assisted by natural or simulated sunlight irradiation, achieving a mineralization degree of 90% at an H2O2 to Ph molar ratio of 6. The authors explained this result based on the high activity of reduced copper to form hydroxyl radicals and the favorable redox process of Fe2+ maintaining a pool of reduced Cu+ species. Thus, a heterogeneous catalyst based on abundant iron or copper metals allowed the promotion of H2O2 activation to yield hydroxyl radicals under visible light irradiation according to Equations (28)–(36).
F e 20 C u 80 / D 3   +   h ν     e +   h +
H 2 O 2 +   e   H O + H O
F e 3 + + e     F e 2 +
C u 2 + + e     C u +
C u + + e     C u 0
h + +     H 2 O 2       H O 2 + H +
h + +     F e 2 +       F e 3 +
h + +     C u +       C u 2 +
h + +     o r g a n i c   c o m p u n d s       o x i d i z e d   p r o d u c t s
Khan et al. [98] compared the photocatalytic activity of novel peculiar–shaped CuO, Fe2O3, and FeO nanoparticles (NPs) to that of the iron(II)–doped copper ferrite, CuII0.4FeII0.6FeIII2O4, through the degradation of MB and RhB. The catalysts were synthesized via the simple co–precipitation and calcination technique. The highest degradation efficiency was achieved by CuO for RhB and by CuII0.4Fe II0.6FeIII2O4 for MB. The CuO/FeO/Fe2O3 composite proved to be the second–best catalyst in both cases, with excellent reusability.
Asenath–Smith et al. [99] used iron oxide (α–Fe2O3, hematite) colloids synthesized under hydrothermal conditions as catalysts for the photodegradation of MO. To enhance the photocatalytic performance, Fe2O3 was combined with other transition–metal oxide (TMO) colloids (e.g., CuO and ZnO), which are sensitive to different regions of the solar spectrum (visible and UV, respectively), using a ternary blending approach for compositional mixtures. For a variety of ZnO/Fe2O3/CuO mole ratios, the pseudo–first–order rate constant for MO degradation was at least twice the sum of the individual Fe2O3 and CuO rate constants, indicating that an underlying synergy governs the photocatalytic reactions for these combinations of TMOs. The increased photo–catalytic performance of Fe2O3 in the presence of CuO was associated with the hydroxyl radical, consistent with heterogeneous photo–Fenton mechanisms, which are not accessible by ZnO. Then, the authors proposed a mechanism where CuO plays a supportive role in Fe2O3 photocatalysis by decomposing H2O2 to generate HO radicals. With additional ROS available, the reaction kinetics are accelerated. The bandgap energies of the TMOs used in this study further corroborated the obtained results. Smaller bandgap materials, such as Fe2O3 and CuO, have valence–band edges above the oxidation potential of H2O. As a result, it is energetically favorable for a photogenerated h+ on Fe2O3 or CuO to participate in the generation of HO radicals by H2O oxidation. Larger band gap materials, such as ZnO, have valence–band edges that are below the H2O oxidation potential for HO formation, but their conduction band edges are below the reduction potential of O2.
Lu et al. [100] used a magnesium aluminum silicate known as attapulgite (ATP) supported Fe–Mn–Cu polymetallic oxide as a catalyst with UV irradiation in the photocatalytic oxidation (photo–Fenton) treatment of a synthetic pharmaceutical wastewater. Fe–Mn–Cu@ATP had good catalytic potential and a significant synergistic effect since it removed almost all heterocyclic compounds, as well as humus–like and fulvic acid. The degradation efficiency of the nanocomposite only decreased by 5.8% after repeated use for six cycles (COD removal reached 64.9%, and the BOD 5/COD increased from 0.179 to 0.387 after 180 min of reaction time). The proposed mechanism involves: (i) the direct reaction between the catalyst and hydrogen peroxide, which generates HO radicals and oxidized ions of the three metals, (ii) the UV photoreductions of these metal ions, and (iii) indirect reactions between photogenerated free radicals and the catalyst. In this mechanism, the consumed metal ions are effectively regenerated under UV light, and HO finally reacts with the organic matter.
Davarnejad et al. [101] synthesized alginate–based hydrogel–coated bimetallic iron–copper nanocomposite beads through a green method and used them as heterogeneous catalysts for metronidazole elimination from wastewater. These authors employed the response surface methodology (RSM) based on the Box–Behnken design (BBD) to assess both the individual and interaction effects of five main variables involving catalyst loading, initial pH, reaction time, metronidazole, and H2O2 concentrations. The data obtained from the model were in good agreement with the experimental ones. The optimum conditions (for 95.3%, based on model, and for 95.0%, based on experiments) were found at a catalyst (Fe2O3–CuO@Ca–Alg) loading of 44.7 mg L−1, an initial pH of 3.5, a metronidazole concentration of 10 mg L−1, a H2O2 concentration of 33.17 mmol L−1, and a reaction time of 85 min.
Zhang et al. [102] used Fe3O4@Cu2O/carbon quantum dots (CQDs)/nitrogen–doped carbon quantum dots (N–CQDs) (FCCN) for the degradation of azo compounds, such as MO, acid orange II, and mordant yellow 10, even at a neutral and alkaline pH (pH: 7–12) with a shortest time for complete degradation of 15 min. There is a cooperative interaction between each component, so the synergism from all the components of FCCN enhances the photocatalytic properties. The authors proposed a mechanism where in the first place, the up–conversion characteristics of CQDs and N–CQDs improve the light utilization of FCCN. Secondly, the load of CQDs and N–CQDs can separate and transfer photo–generated charges between Fe3O4 and Cu2O more efficiently, accelerating the circulation of Fe3+/Fe2+ for the decomposition of H2O2 to yield the active species: HO radical. Thirdly, the light–induced protons of CQDs change the partial pH, so that the FCCN system is able to work in alkaline solutions. Finally, the insoluble layers of CQDs and N–CQDs are beneficial to enhance the stability of the composite (see in Figure 7 the schematic diagram of the catalytic mechanism proposed by the authors). The magnetism of Fe3O4 made this catalyst easily separable and the insoluble layers of CQDs and N–CQDs allowed it to be repeatedly used without activity change even after 10 cycles. The degradation rate constant of MO in the FCCN/H2O2/light system was 5.4 times higher than that of the Fe3O4@Cu2O/H2O2/light system, indicating that the loaded quantum dots greatly enhanced the photocatalytic activity. The degradation rate constant in the FCCN/H2O2/light system was 2.5 times higher than that of the simple mixture (Fe3O4@Cu2O + CQDs + N–CQDs)/H2O2/light system, which suggests that these excellent catalytic performances should come from the synergism effect of all components in FCCN.
All the results described in this section are summarized in Table 3.

6. Metal–Organic Frameworks Based on Fe and Cu

Metal–organic frameworks (MOFs) are inorganic–organic porous polymers composed of metal ions or metal clusters linked to each other by bridging organic ligands to form three–dimensional structures with a high specific surface area. In particular, Fe–based MOFs, made by polydentate organic ligands as linkers and inorganic Fe ions or Fe–oxo clusters as nodes, have recently attracted great attention in various applications such as gas adsorption, catalysis, and sensor development [104]. Within the applications, Fe–based MOFs have been considered promising catalysts in photo–Fenton processes due to their interfacial electron transfer properties and the cycling of the Fe(III)/Fe(II) redox pair. Additionally, the band gaps of Fe–based MOFs are suitable for visible light photoactivation and thus, FeO clusters could be directly excited to generate electron hole pairs (eh+), which subsequently degrade organic pollutants or lead to the generation of reactive oxygen species [105]. Likewise, it has been shown that the separation efficiency of photogenerated carriers in the Fe–O cluster remains low due to fast electron–hole recombination [105,106]. Taking this into account, different modifications in the structure of MOFs have been proposed to delay the recombination rate and improve pollutant degradation efficiencies [106,107]. In this sense, the combination of MOFs with metals or metal oxide nanoparticles may form a new interface or heterojunction that efficiently couples the catalytic process of photo–induced electrons and radicals’ generation [108,109]. Moreover, the addition of a second metal ion into the nodes of frameworks could significantly improve the catalytic properties of MOFs [41]. Recently, different reports have incorporated Cu into Fe–based MOFs structure yielding Fe–Cu bimetallic MOFs in order to improve the performance of these materials in photo–Fenton processes.
Do et al. [110] reported the preparation of Fe–doped Cu 1,4–benzenedicarboxylate MOFs (Fe–CuBDC) and its application as a heterogeneous photo–Fenton catalyst for the degradation of MB in aqueous solution under visible light irradiation. In this work, the authors compared the efficiency of the bimetallic Fe–CuBDC with each single metal MOF. The degradation efficiency of MB with Fe–CuBDC was much higher than those obtained with the others, which indicates that the partial substitution of Cu by Fe significantly improves the photocatalytic activity of the material. Complete removal of MB was achieved after 70 min under light irradiation with 1.0 g L−1 of Fe–CuBDC (Table 4). The degradation performance of the Fe–CuBDC catalyst was relatively constant with a slight reduction in removal efficiency from 99.9% to 97.2% after five reuse cycles.
Shi et al. [111] prepared a binary bimetallic heterojunction, which consisted of a core–shell magnetic CuFe2O4@MIL–100(Fe, Cu) metal–organic framework, via an in–situ derivation strategy. The synthesis methods, which involved the in situ surface pyrolysis of CuFe2O4 nanoparticles and complexation with trimesic acid to derive MIL–100(Fe, Cu), allowed the formation of a bimetallic core–shell CuFe2O4@MIL–100(Fe, Cu) heterojunction (MCuFe MOF). A schematic diagram of the synthesis process of MCuFe MOF is shown in Figure 8. These materials showed notable catalytic performance in a photo–Fenton process towards the degradation of various organic pollutants by increasing H2O2 activation efficiency and decreasing the required dosage of MCuFe MOF (0.05 g L−1) over a wide pH range (4–9) (Table 4). Furthermore, the MCuFe MOF showed a high stability for the degradation of organic contaminants, with almost no decrease in activity and negligible metal leaching after five successive cycles. According to the proposed mechanism, the combination of the photothermal conversion effect with the formed heterojunction cannot only accelerate the generation and separation of photogenerated e and h+ but also improve the continuous and efficient transformation of ≡Fe(III)/Fe(II) and ≡Cu(II)/Cu(I) redox couples, leading to enhanced photo–Fenton efficiency (see Figure 9).
Zhong et al. [112] performed Cu2O growth on the surface of Cu–doped MIL–100 (Fe) to obtain a Cu2O/MIL(Fe/Cu) composite, which showed an enhanced interfacial synergistic effect and was successfully used as photo–Fenton catalysts for thiacloprid (TCL) degradation. The authors reported that Cu doping into MIL(Fe) led to the reduction in the band gap, and a boost of the redox cycle Fe2+/Fe3+. Likewise, the growth of Cu2O extended the light absorption range of MIL(Fe/Cu) from UV to the visible region. A proposed TCL degradation mechanism is shown in Figure 10. Cu2O/MIL(Fe/Cu) composite exhibited a TCL degradation rate nearly 10 times faster than those of Cu2O and MIL–100(Fe), separately, achieving a complete TCL degradation within 20 min of reaction. Moreover, TOC removal reached 82.3% within 80 min under neutral conditions, which highlights the good performance of the Cu2O/MIL(Fe/Cu) composite. Catalyst reuse tests showed no significant efficiency loss in reaction rates even after the tenth cycle, reaching complete TCL degradation within 20 min in each cycle.
Low–crystalline MOFs, with long–range disorder but local crystallinity, allow the availability of more active sites and defects than highly crystalline materials. These materials could improve the activation capacity of Fe–based MOFs towards H2O2 by increasing the number of metallic coordinately unsaturated active sites (CUS) within the frameworks. Wu et al. [113] prepared low–crystalline bimetallic MOFs of MIL–53(Fe, M) (M: Mn or Cu), via a one–pot solvothermal method, as photo–Fenton catalysts for the degradation of CIP. The results showed a significant improvement in photo–Fenton performance for the low–crystallinity catalyst compared to the crystalline counterparts, which was mainly attributed to the enhancement of the synergism between the hetero–metal nodes. In particular, low crystallinity MIL–53(Fe, Cu) exhibited a much higher removal efficiency and a faster reaction rate than that of crystalline MIL–53(Fe, Cu) within 30 min (Table 4). Besides the increased metal CUSs in the low–crystalline state, both Cu and Mn could increase the specific surface area and promote the visible–light absorption and separation/transportation of carriers in the low–crystalline state, thus leading to the acceleration of Fe(II)/Fe(III) and M(red)/M(ox) cycles.
All the results described in this section are summarized in Table 4.

7. Conclusions and Future Perspective

This review has shown the present trends in the development of bimetallic iron–copper materials and their application as catalysts in heterogeneous photo–Fenton–like reactions for the degradation of wastewater pollutants. Bimetallic Fe–Cu catalysts have shown better performance than monometallic Fe catalysts.
We have compared many materials with very different chemical compositions and physicochemical properties. In general, bimetallic Fe–Cu catalysts have shown better performance than monometallic Fe catalysts. In the works summarized here, it is difficult to find a common explanation of the synergistic effect of iron and copper on the catalytic activity of the materials. Some works attribute the improvement to changes in the band–gap, others to the decrease in the electron–hole recombination rates, and even to the reduction of H2O2 by electrons, or the beneficial effect of the binary redox couples of Fe(II)/Fe(III) and Cu(I)/Cu(II) on the decomposition efficiency of H2O2. However, the common factor in all the materials seems to be that the coexistence of both metals evidently favors the redox cycles of Fe and Cu, resulting in higher catalytic activity. The co–existence of Fe and Cu on the surface of a catalyst accelerates the process of electron–transfer in the reaction environment, providing an appropriate condition for the activation of hydrogen peroxide and the generation of reactive radical species. This hypothesis is supported by the excellent degradation and mineralization results obtained when aqueous solutions of the contaminant are treated with these catalysts. In particular, lower degradation times and catalyst dosages are required to achieve complete pollutant degradation. Most of the bimetallic Fe–Cu catalysts have shown very good results at a circumneutral pH, which represents one of the major advantages of these materials since the development of an efficient and sustainable photo–Fenton treatment at neutral pH values remains a challenge for the scientific community working in the field. In addition, many of the works summarized here report the stability of the bimetallic catalysts even after having been used in four or five cycles.
From the analysis of the collected research, we can conclude that further efforts should focus on the preparation of new catalysts by green synthesis routes, that is, using bioactive agents such as plant materials, microorganisms, or biological waste. This approach would help to reduce the overall production cost and to limit environmental impact. Only a few of the compilated articles employed renewable energies such as natural solar light as the photoirradiation source. Most of them used either Xe lamps with or without cut–off filters for simulating sunlight or other light sources, such as UV lamps or visible light from lamps or LEDs. In this context, further studies devoted to addressing the effect of using solar light as the photoirradiation source on the performance of the catalysts are certainly encouraged.
Although excellent degradation and mineralization results have been obtained by applying this technique at the laboratory scale, its effectiveness needs to be evaluated in continuous flow systems for real industrial effluents. Most of the published research focused on the degradation of single pollutants (mainly dyes and antibiotics), disregarding the effects of the presence of humic–like dissolved organic matter and other contaminants. Moreover, to increase the efficiency and decrease the cost of the treatment of recalcitrant contaminants in real systems, the integration of photo–Fenton methods with biological technologies should be explored.

Author Contributions

Conceptualization, investigation, writing—original draft preparation, writing—review and editing: G.N.B., F.S.G.E., L.C. and D.O.M. contributed equally. All authors have read and agreed to the published version of the manuscript.

Funding

This project received funding from ANPCyT, Argentina (PICT 2017–1628, and PICT 2019–03140).

Data Availability Statement

Not applicable.

Acknowledgments

G.N. Bosio, F. García Einschlag, and L. Carlos and are staff researchers of CONICET (Argentina). D.O. Mártire is staff researcher of Comisión de Investigaciones Científicas (CIC, Buenos Aires, Argentina).

Conflicts of Interest

There are no conflicts of competing interest to declare.

References

  1. Afrad, M.S.I.; Monir, M.B.; Haque, M.E.; Barau, A.A.; Haque, M.M. Impact of Industrial Effluent on Water, Soil and Rice Production in Bangladesh: A Case of Turag River Bank. J. Environ. Health Sci. Eng. 2020, 18, 825–834. [Google Scholar] [CrossRef] [PubMed]
  2. Amor, C.; Marchão, L.; Lucas, M.S.; Peres, J.A. Application of Advanced Oxidation Processes for the Treatment of Recalcitrant Agro–Industrial Wastewater: A Review. Water 2019, 11, 205. [Google Scholar] [CrossRef] [Green Version]
  3. Lama, G.; Meijide, J.; Sanromán, A.; Pazos, M. Heterogeneous Advanced Oxidation Processes: Current Approaches for Wastewater Treatment. Catalysts 2022, 12, 344. [Google Scholar] [CrossRef]
  4. Shokri, A.; Fard, M.S. A Critical Review in Fenton–like Approach for the Removal of Pollutants in the Aqueous Environment. Environ. Chall. 2022, 7, 100534. [Google Scholar] [CrossRef]
  5. Aparicio, F.; Escalada, J.P.; De Gerónimo, E.; Aparicio, V.C.; Einschlag, F.S.G.; Magnacca, G.; Carlos, L.; Mártire, D.O. Carbamazepine Degradation Mediated by Light in the Presence of Humic Substances–Coated Magnetite Nanoparticles. Nanomaterials 2019, 9, 1379. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  6. Litter, M.I.; Slodowicz, M. An Overview on Heterogeneous Fenton and PhotoFenton Reactions Using Zerovalent Iron Materials. J. Adv. Oxid. Technol. 2017, 20, 20160164. [Google Scholar] [CrossRef]
  7. 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]
  8. 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]
  9. Hussain, S.; Aneggi, E.; Goi, D. Catalytic Activity of Metals in Heterogeneous Fenton–like Oxidation of Wastewater Contaminants: A Review. Environ. Chem. Lett. 2021, 19, 2405–2424. [Google Scholar] [CrossRef]
  10. Nichela, D.A.; Berkovic, A.M.; Costante, M.R.; Juliarena, M.P.; García Einschlag, F.S. Nitrobenzene Degradation in Fenton–like Systems Using Cu(II) as Catalyst. Comparison between Cu(II)– and Fe(III)–Based Systems. Chem. Eng. J. 2013, 228, 1148–1157. [Google Scholar] [CrossRef]
  11. Berkovic, A.M.; Costante, M.R.; García Einschlag, F.S. Combining Multivariate Curve Resolution and Lumped Kinetic Modelling for the Analysis of Lignin Degradation by Copper–Catalyzed Fenton–like Systems. React. Chem. Eng. 2022, 7, 1954–1967. [Google Scholar] [CrossRef]
  12. Bali, U.; Karagözoǧlu, B. Performance Comparison of Fenton Process, Ferric Coagulation and H2O2/Pyridine/Cu(II) System for Decolorization of Remazol Turquoise Blue G–133. Dye Pigment. 2007, 74, 73–80. [Google Scholar] [CrossRef]
  13. Tian, X.; Jin, H.; Nie, Y.; Zhou, Z.; Yang, C.; Li, Y.; Wang, Y. Heterogeneous Fenton–like Degradation of Ofloxacin over a Wide PH Range of 3.6–10.0 over Modified Mesoporous Iron Oxide. Chem. Eng. J. 2017, 328, 397–405. [Google Scholar] [CrossRef]
  14. Do, Q.C.; Kim, D.G.; Ko, S.O. Catalytic Activity Enhancement of a Fe3O4@SiO2 Yolk–Shell Structure for Oxidative Degradation of Acetaminophen by Decoration with Copper. J. Clean. Prod. 2018, 172, 1243–1253. [Google Scholar] [CrossRef]
  15. Salem, I.A. Kinetics of the Oxidative Color Removal and Degradation of Bromophenol Blue with Hydrogen Peroxide Catalyzed by Copper(II)–Supported Alumina and Zirconia. Appl. Catal. B Environ. 2000, 28, 153–162. [Google Scholar] [CrossRef]
  16. Pignatello, J.; Oliveros, E.; MacKay, A. Advanced oxidation processes for organic contaminant destruction based on the fenton reaction and related chemistry. Crit. Rev. Environ. Sci. Technol. 2006, 36, 1–84. [Google Scholar] [CrossRef]
  17. Zhang, M.H.; Dong, H.; Zhao, L.; Wang, D.-X.; Meng, D. A Review on Fenton Process for Organic Wastewater Treatment Based on Optimization Perspective. Sci. Total Environ. 2019, 670, 110–121. [Google Scholar] [CrossRef]
  18. McNaught, A.D.; Wilkinson, A. The IUPAC Compendium of Chemical Terminology; International Union of Pure and Applied Chemistry (IUPAC): Durham, NC, USA, 2019; ISBN 0-9678550-9-8. [Google Scholar]
  19. Sun, C.; Chen, C.; Ma, W.; Zhao, J. Photodegradation of Organic Pollutants Catalyzed by Iron Species under Visible Light Irradiation. Phys. Chem. Chem. Phys. 2011, 13, 1957–1969. [Google Scholar] [CrossRef] [PubMed]
  20. Kalal, S.; Singh Chauhan, N.P.; Ameta, N.; Ameta, R.; Kumar, S.; Punjabi, P.B. Role of Copper Pyrovanadate as Heterogeneous Photo–Fenton like Catalyst for the Degradation of Neutral Red and Azure–B: An Eco–Friendly Approach. Korean J. Chem. Eng. 2014, 31, 2183–2191. [Google Scholar] [CrossRef]
  21. Lee, H.J.; Lee, H.; Lee, C. Degradation of Diclofenac and Carbamazepine by the Copper(II)–Catalyzed Dark and Photo–Assisted Fenton–like Systems. Chem. Eng. J. 2014, 245, 258–264. [Google Scholar] [CrossRef]
  22. Faust, B.C.; Hoigné, J. Photolysis of Fe (III)–Hydroxy Complexes as Sources of OH Radicals in Clouds, Fog and Rain. Atmos. Environ. Part A Gen. Top. 1990, 24, 79–89. [Google Scholar] [CrossRef]
  23. Malato, S.; Fernández–Ibáñez, P.; Maldonado, M.I.; Blanco, J.; Gernjak, W. Decontamination and Disinfection of Water by Solar Photocatalysis: Recent Overview and Trends. Catal. Today 2009, 147, 1–59. [Google Scholar] [CrossRef]
  24. García Einschlag, F.S.; Braun, A.M.; Oliveros, E. Fundamentals and Applications of the Photo–Fenton Process to Water Treatment. In Handbook of Environmental Chemistry; Springer: Berlin/Heidelberg, Germany, 2015; Volume 35, pp. 301–342. [Google Scholar]
  25. Bauer, R.; Fallmann, H. The Photo–Fenton Oxidation—A Cheap and Efficient Wastewater Treatment Method. Res. Chem. Intermed. 1997, 23, 341–354. [Google Scholar] [CrossRef]
  26. Pliego, G.; Zazo, J.A.; Garcia–Muñoz, P.; Munoz, M.; Casas, J.A.; Rodriguez, J.J. Trends in the Intensification of the Fenton Process for Wastewater Treatment: An Overview. Crit. Rev. Environ. Sci. Technol. 2015, 45, 2611–2692. [Google Scholar] [CrossRef]
  27. Rahim Pouran, S.; Abdul Aziz, A.R.; Wan Daud, W.M.A. Review on the Main Advances in Photo–Fenton Oxidation System for Recalcitrant Wastewaters. J. Ind. Eng. Chem. 2015, 21, 53–69. [Google Scholar] [CrossRef]
  28. Miralles–Cuevas, S.; Oller, I.; Pérez, J.A.S.; Malato, S. Application of Solar Photo–Fenton at Circumneutral PH to Nanofiltration Concentrates for Removal of Pharmaceuticals in MWTP Effluents. Environ. Sci. Pollut. Res. 2015, 22, 846–855. [Google Scholar] [CrossRef]
  29. Maniakova, G.; Salmerón, I.; Aliste, M.; Inmaculada Polo–López, M.; Oller, I.; Malato, S.; Rizzo, L. Solar Photo–Fenton at Circumneutral PH Using Fe(III)–EDDS Compared to Ozonation for Tertiary Treatment of Urban Wastewater: Contaminants of Emerging Concern Removal and Toxicity Assessment. Chem. Eng. J. 2022, 431, 133474. [Google Scholar] [CrossRef]
  30. López–Vinent, N.; Cruz–Alcalde, A.; Lai, C.; Giménez, J.; Esplugas, S.; Sans, C. Role of Sunlight and Oxygen on the Performance of Photo–Fenton Process at near Neutral PH Using Organic Fertilizers as Iron Chelates. Sci. Total Environ. 2022, 803, 149873. [Google Scholar] [CrossRef]
  31. Ahile, U.J.; Wuana, R.A.; Itodo, A.U.; Sha’Ato, R.; Dantas, R.F. A Review on the Use of Chelating Agents as an Alternative to Promote Photo–Fenton at Neutral PH: Current Trends, Knowledge Gap and Future Studies. Sci. Total Environ. 2020, 710, 134872. [Google Scholar] [CrossRef]
  32. Sanabria, P.; Wilde, M.L.; Ruiz–Padillo, A.; Sirtori, C. Trends in Fenton and Photo–Fenton Processes for Degradation of Antineoplastic Agents in Water Matrices: Current Knowledge and Future Challenges Evaluation Using a Bibliometric and Systematic Analysis. Environ. Sci. Pollut. Res. 2022, 29, 42168–42184. [Google Scholar] [CrossRef]
  33. Matta, R.; Hanna, K.; Kone, T.; Chiron, S. Oxidation of 2,4,6-trinitrotoluene in the presence of different iron-bearing minerals at neutral pH. Chem. Eng. J. 2008, 144, 453–458. [Google Scholar] [CrossRef]
  34. 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] [PubMed]
  35. Zuo, S.; Jin, X.; Wang, X.; Lu, Y.; Zhu, Q.; Wang, J.; Liu, W.; Du, Y.; Wang, J. Sandwich Structure Stabilized Atomic Fe Catalyst for Highly Efficient Fenton–like Reaction at All PH Values. Appl. Catal. B Environ. 2021, 282, 119551. [Google Scholar] [CrossRef]
  36. Xu, L.; Wang, J. Degradation of 2,4,6–Trichlorophenol Using Magnetic Nanoscaled Fe3O4/CeO2 Composite as a Heterogeneous Fenton–like Catalyst. Sep. Purif. Technol. 2015, 149, 255–264. [Google Scholar] [CrossRef]
  37. Wang, J.; Tang, J. Fe–Based Fenton–like Catalysts for Water Treatment: Catalytic Mechanisms and Applications. J. Mol. Liq. 2021, 332, 115755. [Google Scholar] [CrossRef]
  38. Xin, S.; Ma, B.; Liu, G.; Ma, X.; Zhang, C.; Ma, X.; Gao, M.; Xin, Y. Enhanced Heterogeneous Photo–Fenton–like Degradation of Tetracycline over CuFeO2/Biochar Catalyst through Accelerating Electron Transfer under Visible Light. J. Environ. Manag. 2021, 285, 112093. [Google Scholar] [CrossRef]
  39. Wei, Y.; Wang, C.; Liu, D.; Jiang, L.; Chen, X.; Li, H.; Zhang, F. Photo–Catalytic Oxidation for Pyridine in Circumneutral Aqueous Solution by Magnetic Fe–Cu Materials Activated H2O2. Chem. Eng. Res. Des. 2020, 163, 1–11. [Google Scholar] [CrossRef]
  40. Guo, X.; Xu, Y.; Wang, K.; Zha, F.; Tang, X.; Tian, H. Synthesis of Magnetic CuFe2O4 Self–Assembled Hollow Nanospheres and Its Application for Degrading Methylene Blue. Res. Chem. Intermed. 2020, 46, 853–869. [Google Scholar] [CrossRef]
  41. Tang, J.; Wang, J. Iron–Copper Bimetallic Metal–Organic Frameworks for Efficient Fenton–like Degradation of Sulfamethoxazole under Mild Conditions. Chemosphere 2020, 241, 125002. [Google Scholar] [CrossRef]
  42. Han, Z.; Dong, Y.; Dong, S. Copper–Iron Bimetal Modified PAN Fiber Complexes as Novel Heterogeneous Fenton Catalysts for Degradation of Organic Dye under Visible Light Irradiation. J. Hazard. Mater. 2011, 189, 241–248. [Google Scholar] [CrossRef]
  43. Sun, Y.; Yang, Z.; Tian, P.; Sheng, Y.; Xu, J.; Han, Y.F. Oxidative Degradation of Nitrobenzene by a Fenton–like Reaction with Fe–Cu Bimetallic Catalysts. Appl. Catal. B Environ. 2019, 244, 1–10. [Google Scholar] [CrossRef]
  44. Diodati, S.; Walton, R.I.; Mascotto, S.; Gross, S. Low–Temperature Wet Chemistry Synthetic Approaches towards Ferrites. Inorg. Chem. Front. 2020, 7, 3282–3314. [Google Scholar] [CrossRef]
  45. Bastianello, M.; Gross, S.; Elm, M.T. Thermal Stability, Electrochemical and Structural Characterization of Hydrothermally Synthesised Cobalt Ferrite (CoFe2O4). RSC Adv. 2019, 9, 33282–33289. [Google Scholar] [CrossRef] [Green Version]
  46. Ranga, R.; Kumar, A.; Kumari, P.; Singh, P.; Madaan, V.; Kumar, K. Ferrite Application as an Electrochemical Sensor: A Review. Mater. Charact. 2021, 178, 111269. [Google Scholar] [CrossRef]
  47. Rana, G.; Dhiman, P.; Kumar, A.; Vo, D.V.N.; Sharma, G.; Sharma, S.; Naushad, M. Recent Advances on Nickel Nano–Ferrite: A Review on Processing Techniques, Properties and Diverse Applications. Chem. Eng. Res. Des. 2021, 175, 182–208. [Google Scholar] [CrossRef]
  48. Kaur, B.; Kaushal, G.; Rana, S.; Kumar, P.; Khanra, P.; Dhiman, M. Magnetic Ferrites: A Brief Review About Substation on Electric and Magnetic Properties. ECS Trans. 2022, 107, 9093–9101. [Google Scholar] [CrossRef]
  49. Almessiere, M.A.; Slimani, Y.; Trukhanov, A.V.; Sadaqat, A.; Korkmaz, A.D.; Algarou, N.A.; Aydın, H.; Baykal, A.; Toprak, M.S. Review on Functional Bi–Component Nanocomposites Based on Hard/Soft Ferrites: Structural, Magnetic, Electrical and Microwave Absorption Properties. Nano-Struct. Nano-Objects 2021, 26, 100728. [Google Scholar] [CrossRef]
  50. Chand, P.; Vaish, S.; Kumar, P. Structural, Optical and Dielectric Properties of Transition Metal (MFe2O4; M = Co, Ni and Zn) Nanoferrites. Phys. B Condens. Matter 2017, 524, 53–63. [Google Scholar] [CrossRef]
  51. Kefeni, K.K.; Mamba, B.B. Photocatalytic Application of Spinel Ferrite Nanoparticles and Nanocomposites in Wastewater Treatment: Review. Sustain. Mater. Technol. 2020, 23, e00140. [Google Scholar] [CrossRef]
  52. Shobana, M.K. Nanoferrites in Biosensors—A Review. Mater. Sci. Eng. B Solid-State Mater. Adv. Technol. 2021, 272, 115344. [Google Scholar] [CrossRef]
  53. Amiri, M.; Salavati–Niasari, M.; Akbari, A. Magnetic Nanocarriers: Evolution of Spinel Ferrites for Medical Applications. Adv. Colloid Interface Sci. 2019, 265, 29–44. [Google Scholar] [CrossRef] [PubMed]
  54. Alao, A.O.; Popoola, A.P.; Sanni, O. The Influence of Nanoparticle Inhibitors on the Corrosion Protection of Some Industrial Metals: A Review. J. Bio- Tribo-Corrosion 2022, 8, 68. [Google Scholar] [CrossRef]
  55. Mmelesi, O.K.; Masunga, N.; Kuvarega, A.; Nkambule, T.T.; Mamba, B.B.; Kefeni, K.K. Cobalt Ferrite Nanoparticles and Nanocomposites: Photocatalytic, Antimicrobial Activity and Toxicity in Water Treatment. Mater. Sci. Semicond. Process. 2021, 123, 105523. [Google Scholar] [CrossRef]
  56. Dippong, T.; Levei, E.A.; Cadar, O. Recent Advances in Synthesis and Applications of MFe2O4 (M = Co, Cu, Mn, Ni, Zn) Nanoparticles. Nanomaterials 2021, 11, 1560. [Google Scholar] [CrossRef]
  57. Iqbal, M.J.; Yaqub, N.; Sepiol, B.; Ismail, B. A Study of the Influence of Crystallite Size on the Electrical and Magnetic Properties of CuFe2O4. Mater. Res. Bull. 2011, 46, 1837–1842. [Google Scholar] [CrossRef]
  58. Bagade, A.; Nagwade, P.; Nagawade, A.; Thopate, S.; Pandit, V.; Pund, S. Impact of Mg2+ Substitution on Structural, Magnetic and Optical Properties of Cu–Cd Ferrites. Mater. Today Proc. 2022, 53, 144–152. [Google Scholar] [CrossRef]
  59. Holinsworth, B.S.; Mazumdar, D.; Sims, H.; Sun, Q.C.; Yurtisigi, M.K.; Sarker, S.K.; Gupta, A.; Butler, W.H.; Musfeldt, J.L. Chemical Tuning of the Optical Band Gap in Spinel Ferrites: CoFe2O4 vs. NiFe2O4. Appl. Phys. Lett. 2013, 103, 082406. [Google Scholar] [CrossRef]
  60. Lai, Y.J.; Lee, D.J. Solid Mediator Z–Scheme Heterojunction Photocatalysis for Pollutant Oxidation in Water: Principles and Synthesis Perspectives. J. Taiwan Inst. Chem. Eng. 2021, 125, 88–114. [Google Scholar] [CrossRef]
  61. Feng, Y.; Lee, P.H.; Wu, D.; Zhou, Z.; Li, H.; Shih, K. Degradation of Contaminants by Cu+–Activated Molecular Oxygen in Aqueous Solutions: Evidence for Cupryl Species (Cu3+). J. Hazard. Mater. 2017, 331, 81–87. [Google Scholar] [CrossRef]
  62. Silva, E.D.N.; Brasileiro, I.L.O.; Madeira, V.S.; De Farias, B.A.; Ramalho, M.L.A.; Rodríguez–Aguado, E.; Rodríguez–Castellón, E. Reusable CuFe2O4–Fe2O3 catalyst Synthesis and Application for the Heterogeneous Photo–Fenton Degradation of Methylene Blue in Visible Light. J. Environ. Chem. Eng. 2020, 8, 104132. [Google Scholar] [CrossRef]
  63. Cao, Z.; Zuo, C.; Wu, H. One Step for Synthesis of Magnetic CuFe2O4 Composites as Photo–Fenton Catalyst for Degradation Organics. Mater. Chem. Phys. 2019, 237, 121842. [Google Scholar] [CrossRef]
  64. Leichtweis, J.; Silvestri, S.; Welter, N.; Vieira, Y.; Zaragoza–Sánchez, P.I.; Chávez–Mejía, A.C.; Carissimi, E. Wastewater Containing Emerging Contaminants Treated by Residues from the Brewing Industry Based on Biochar as a New CuFe2O4/Biochar Photocatalyst. Process Saf. Environ. Prot. 2021, 150, 497–509. [Google Scholar] [CrossRef]
  65. Jiang, J.; Gao, J.; Niu, S.; Wang, X.; Li, T.; Liu, S.; Lin, Y.; Xie, T.; Dong, S. Comparing Dark– and Photo–Fenton–like Degradation of Emerging Pollutant over Photo–Switchable Bi2WO6/CuFe2O4: Investigation on Dominant Reactive Oxidation Species. J. Environ. Sci. 2021, 106, 147–160. [Google Scholar] [CrossRef]
  66. Clarizia, L.; Russo, D.; Di Somma, I.; Marotta, R.; Andreozzi, R. Homogeneous Photo–Fenton Processes at near Neutral PH: A Review. Appl. Catal. B Environ. 2017, 209, 358–371. [Google Scholar] [CrossRef]
  67. Guo, X.; Wang, K.; Xu, Y. Tartaric Acid Enhanced CuFe2O4–Catalyzed Heterogeneous Photo–Fenton–like Degradation of Methylene Blue. Mater. Sci. Eng. B Solid-State Mater. Adv. Technol. 2019, 245, 75–84. [Google Scholar] [CrossRef]
  68. Li, Y.; Qin, C.; Zhang, J.; Lan, Y.; Zhou, L. Cu(II) Catalytic Reduction of Cr(VI) by Tartaric Acid under the Irradiation of Simulated Solar Light. Environ. Eng. Sci. 2014, 31, 447–452. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  69. Rocha, A.K.S.; Magnago, L.B.; Santos, J.J.; Leal, V.M.; Marins, A.A.L.; Pegoretti, V.C.B.; Ferreira, S.A.D.; Lelis, M.F.F.; Freitas, M.B.J.G. Copper Ferrite Synthesis from Spent Li–Ion Batteries for Multifunctional Application as Catalyst in Photo Fenton Process and as Electrochemical Pseudocapacitor. Mater. Res. Bull. 2019, 113, 231–240. [Google Scholar] [CrossRef]
  70. Lin, Y.Y.; Lu, S.Y. Selective and Efficient Cleavage of Lignin Model Compound into Value–Added Aromatic Chemicals with CuFe2O4 Nanoparticles Decorated on Partially Reduced Graphene Oxides via Sunlight–Assisted Heterogeneous Fenton Processes. J. Taiwan Inst. Chem. Eng. 2019, 97, 264–271. [Google Scholar] [CrossRef]
  71. Ayala, L.I.M.; Paquet, M.; Janowska, K.; Jamard, P.; Quist–Jensen, C.A.; Bosio, G.N.; Mártire, D.O.; Fabbri, D.; Boffa, V. Water Defluoridation: Nanofiltration vs. Membrane Distillation. Ind. Eng. Chem. Res. 2018, 57, 14740–14748. [Google Scholar] [CrossRef] [Green Version]
  72. Wang, T.; Wang, Z.; Wang, P.; Tang, Y. An Integration of Photo–Fenton and Membrane Process for Water Treatment by a PVDF@CuFe2O4 Catalytic Membrane. J. Membr. Sci. 2019, 572, 419–427. [Google Scholar] [CrossRef]
  73. Guin, D.; Baruwati, B.; Manorama, S.V. A Simple Chemical Synthesis of Nanocrystalline AFe2O4 (A = Fe, Ni, Zn): An Efficient Catalyst for Selective Oxidation of Styrene. J. Mol. Catal. A Chem. 2005, 242, 26–31. [Google Scholar] [CrossRef]
  74. Zhenmin, L.; Xiaoyong, L.; Hong, W.; Dan, M.; Chaojian, X.; Dan, W. General Synthesis of Homogeneous Hollow Core–Shell Ferrite Microspheres. J. Phys. Chem. C 2009, 113, 2792–2797. [Google Scholar] [CrossRef]
  75. Sharma, R.; Bansal, S.; Singhal, S. Tailoring the Photo–Fenton Activity of Spinel Ferrites (MFe2O4) by Incorporating Different Cations (M = Cu, Zn, Ni and Co) in the Structure. RSC Adv. 2015, 5, 6006–6018. [Google Scholar] [CrossRef]
  76. Li, N.; Fu, F.; Lu, J.; Ding, Z.; Tang, B.; Pang, J. Facile Preparation of Magnetic Mesoporous MnFe2O4@SiO2−CTAB Composites for Cr(VI) Adsorption and Reduction. Environ. Pollut. 2017, 220, 1376–1385. [Google Scholar] [CrossRef]
  77. Deng, J.; Xu, M.; Qiu, C.; Chen, Y.; Ma, X.; Gao, N.; Li, X. Magnetic MnFe2O4 Activated Peroxymonosulfate Processes for Degradation of Bisphenol A: Performance, Mechanism and Application Feasibility. Appl. Surf. Sci. 2018, 459, 138–147. [Google Scholar] [CrossRef]
  78. Velinov, N.; Petrova, T.; Genova, I.; Ivanov, I.; Tsoncheva, T.; Idakiev, V.; Kunev, B.; Mitov, I. Synthesis and Mössbauer Spectroscopic Investigation of Copper–Manganese Ferrite Catalysts for Water–Gas Shift Reaction and Methanol Decomposition. Mater. Res. Bull. 2017, 95, 556–562. [Google Scholar] [CrossRef]
  79. Fu, W.; Yi, J.; Cheng, M.; Liu, Y.; Zhang, G.; Li, L.; Du, L.; Li, B.; Wang, G.; Yang, X. When Bimetallic Oxides and Their Complexes Meet Fenton–like Process. J. Hazard. Mater. 2022, 424, 127419. [Google Scholar] [CrossRef] [PubMed]
  80. Meena, S.; Anantharaju, K.S.; Vidya, Y.S.; Renuka, L.; Uma, B.; Sharma, S.C.; Prasad, B.D.; More, S.S. Enhanced Sunlight Driven Photocatalytic Activity and Electrochemical Sensing Properties of Ce–Doped MnFe2O4 Nano Magnetic Ferrites. Ceram. Int. 2021, 47, 14760–14774. [Google Scholar] [CrossRef]
  81. Niu, J.; Qian, H.; Liu, J.; Liu, H.; Zhang, P.; Duan, E. Process and Mechanism of Toluene Oxidation Using Cu1–YMn2CeyOx/Sepiolite Prepared by the Co–Precipitation Method. J. Hazard. Mater. 2018, 357, 332–340. [Google Scholar] [CrossRef] [PubMed]
  82. Sun, Y.; Zhou, J.; Liu, D.; Li, X.; Liang, H. Enhanced Catalytic Performance of Cu–Doped MnFe2O4 Magnetic Ferrites: Tetracycline Hydrochloride Attacked by Superoxide Radicals Efficiently in a Strong Alkaline Environment. Chemosphere 2022, 297, 134154. [Google Scholar] [CrossRef]
  83. Yang, J.; Zhang, Y.; Zeng, D.; Zhang, B.; Hassan, M.; Li, P.; Qi, C.; He, Y. Enhanced Catalytic Activation of Photo–Fenton Process by Cu0·5Mn0·5Fe2O4 for Effective Removal of Organic Contaminants. Chemosphere 2020, 247, 125780. [Google Scholar] [CrossRef] [PubMed]
  84. Wang, X.T.; Li, Y.; Zhang, X.Q.; Li, J.F.; Luo, Y.N.; Wang, C.W. Fabrication of a Magnetically Separable Cu2ZnSnS4/ZnFe2O4 p–n Heterostructured Nano–Photocatalyst for Synergistic Enhancement of Photocatalytic Activity Combining with Photo–Fenton Reaction. Appl. Surf. Sci. 2019, 479, 86–95. [Google Scholar] [CrossRef]
  85. Shi, L.; Shi, Y.; Zhuo, S.; Zhang, C.; Aldrees, Y.; Aleid, S.; Wang, P. Multi–Functional 3D Honeycomb Ceramic Plate for Clean Water Production by Heterogeneous Photo–Fenton Reaction and Solar–Driven Water Evaporation. Nano Energy 2019, 60, 222–230. [Google Scholar] [CrossRef]
  86. Schmachtenberg, N.; Silvestri, S.; Da Silveira Salla, J.; Dotto, G.L.; Hotza, D.; Jahn, S.L.; Foletto, E.L. Preparation of Delafossite–Type CuFeO2 Powders by Conventional and Microwave–Assisted Hydrothermal Routes for Use as Photo–Fenton Catalysts. J. Environ. Chem. Eng. 2019, 7, 102954. [Google Scholar] [CrossRef]
  87. Liu, Q.L.; Zhao, Z.Y.; Zhao, R.D.; Yi, J.H. Fundamental Properties of Delafossite CuFeO2 as Photocatalyst for Solar Energy Conversion. J. Alloys Compd. 2020, 819, 153032. [Google Scholar] [CrossRef]
  88. da Silveira Salla, J.; da Boit Martinello, K.; Dotto, G.L.; García–Díaz, E.; Javed, H.; Alvarez, P.J.J.; Foletto, E.L. Synthesis of Citrate–Modified CuFeS2 Catalyst with Significant Effect on the Photo–Fenton Degradation Efficiency of Bisphenol a under Visible Light and near–Neutral PH. Colloids Surfaces A Physicochem. Eng. Asp. 2020, 595, 124679. [Google Scholar] [CrossRef]
  89. Da Silveira Salla, J.; Dotto, G.L.; Hotza, D.; Landers, R.; Da Boit Martinello, K.; Foletto, E.L. Enhanced Catalytic Performance of CuFeS2 chalcogenide Prepared by Microwave–Assisted Route for Photo–Fenton Oxidation of Emerging Pollutant in Water. J. Environ. Chem. Eng. 2020, 8, 104077. [Google Scholar] [CrossRef]
  90. Cai, X.; Huang, Q.; Hong, Z.; Zhang, Y.; Hu, H.; Huang, Z.; Liang, J.; Qin, Y. Cu Anchored on Manganese Residue through Mechanical Activation to Prepare a Fe–Cu@SiO2/Starch–Derived Carbon Composites with Highly Stable and Active Visible Light Photocatalytic Performance. J. Environ. Chem. Eng. 2021, 9, 104710. [Google Scholar] [CrossRef]
  91. Xin, S.; Huo, S.; Zhang, C.; Ma, X.; Liu, W.; Xin, Y.; Gao, M. Coupling Nitrogen/Oxygen Self–Doped Biomass Porous Carbon Cathode Catalyst with CuFeO2/Biochar Particle Catalyst for the Heterogeneous Visible–Light Driven Photo–Electro–Fenton Degradation of Tetracycline. Appl. Catal. B Environ. 2022, 305, 121024. [Google Scholar] [CrossRef]
  92. Xin, S.; Huo, S.; Xin, Y.; Gao, M.; Wang, Y.; Liu, W.; Zhang, C.; Ma, X. Heterogeneous Photo–Electro–Fenton Degradation of Tetracycline through Nitrogen/Oxygen Self–Doped Porous Biochar Supported CuFeO2 Multifunctional Cathode Catalyst under Visible Light. Appl. Catal. B Environ. 2022, 312, 121442. [Google Scholar] [CrossRef]
  93. Aparicio, F.; Mizrahi, M.; Ramallo–López, J.M.; Laurenti, E.; Magnacca, G.; Carlos, L.; Mártire, D.O. Novel Bimetallic Magnetic Nanocomposites Obtained from Waste–Sourced Bio–Based Substances as Sustainable Photocatalysts. Mater. Res. Bull. 2022, 152, 111846. [Google Scholar] [CrossRef]
  94. Ayala, L.I.M.; Aparicio, F.; Boffa, V.; Magnacca, G.; Carlos, L.; Bosio, G.N.; Mártire, D.O. Removal of As(III) via Adsorption and Photocatalytic Oxidation with Magnetic Fe–Cu Nanocomposites. Photochem. Photobiol. Sci. 2022, 1, 1–10. [Google Scholar] [CrossRef]
  95. Mansoori, S.; Davarnejad, R.; Ozumchelouei, E.J.; Ismail, A.F. Activated Biochar Supported Iron–Copper Oxide Bimetallic Catalyst for Degradation of Ciprofloxacin via Photo–Assisted Electro–Fenton Process: A Mild PH Condition. J. Water Process Eng. 2021, 39, 101888. [Google Scholar] [CrossRef]
  96. Espinosa, J.C.; Catalá, C.; Navalón, S.; Ferrer, B.; Álvaro, M.; García, H. Iron Oxide Nanoparticles Supported on Diamond Nanoparticles as Efficient and Stable Catalyst for the Visible Light Assisted Fenton Reaction. Appl. Catal. B Environ. 2018, 226, 242–251. [Google Scholar] [CrossRef]
  97. Manickam–Periyaraman, P.; Espinosa, J.C.; Ferrer, B.; Subramanian, S.; Álvaro, M.; García, H.; Navalón, S. Bimetallic Iron–Copper Oxide Nanoparticles Supported on Nanometric Diamond as Efficient and Stable Sunlight–Assisted Fenton Photocatalyst. Chem. Eng. J. 2020, 393, 124770. [Google Scholar] [CrossRef]
  98. Khan, A.; Valicsek, Z.; Horváth, O. Comparing the Degradation Potential of Copper(II), Iron(II), Iron(III) Oxides, and Their Composite Nanoparticles in a Heterogeneous Photo–Fenton System. Nanomaterials 2021, 11, 225. [Google Scholar] [CrossRef]
  99. Asenath–Smith, E.; Ambrogi, E.K.; Barnes, E.; Brame, J.A. CuO Enhances the Photocatalytic Activity of Fe2O3 through Synergistic Reactive Oxygen Species Interactions. Colloids Surfaces A Physicochem. Eng. Asp. 2020, 603, 125179. [Google Scholar] [CrossRef]
  100. Lu, M.; Wang, J.; Wang, Y.; He, Z. Heterogeneous Photo–Fenton Catalytic Degradation of Practical Pharmaceutical Wastewater by Modified Attapulgite Supported Multi–Metal Oxides. Water 2021, 13, 156. [Google Scholar] [CrossRef]
  101. Davarnejad, R.; Hassanvand, Z.R.; Mansoori, S.; Kennedy, J.F. Metronidazole Elimination from Wastewater through Photo–Fenton Process Using Green–Synthesized Alginate–Based Hydrogel Coated Bimetallic Iron-copper Nanocomposite Beads as a Reusable Heterogeneous Catalyst. Bioresour. Technol. Rep. 2022, 18, 101068. [Google Scholar] [CrossRef]
  102. Zhang, Y.; Guo, P.; Jin, M.; Gao, G.; Xi, Q.; Zhou, H.; Xu, G.; Xu, J. Promoting the Photo–Fenton Catalytic Activity with Carbon Dots: Broadening Light Absorption, Higher Applicable PH and Better Reuse Performance. Mol. Catal. 2020, 481, 110254. [Google Scholar] [CrossRef]
  103. Zhang, B.; Hou, Y.; Yu, Z.; Liu, Y.; Huang, J.; Qian, L.; Xiong, J. Three–Dimensional Electro–Fenton Degradation of Rhodamine B with Efficient Fe–Cu/Kaolin Particle Electrodes: Electrodes Optimization, Kinetics, Influencing Factors and Mechanism. Sep. Purif. Technol. 2019, 210, 60–68. [Google Scholar] [CrossRef]
  104. Joseph, J.; Iftekhar, S.; Srivastava, V.; Fallah, Z.; Zare, E.N.; Sillanpää, M. Iron–Based Metal–Organic Framework: Synthesis, Structure and Current Technologies for Water Reclamation with Deep Insight into Framework Integrity. Chemosphere 2021, 284, 131171. [Google Scholar] [CrossRef] [PubMed]
  105. Wang, D.; Wang, M.; Li, Z. Fe–Based Metal–Organic Frameworks for Highly Selective Photocatalytic Benzene Hydroxylation to Phenol. ACS Catal. 2015, 5, 6852–6857. [Google Scholar] [CrossRef]
  106. Ahmad, M.; Chen, S.; Ye, F.; Quan, X.; Afzal, S.; Yu, H.; Zhao, X. Efficient Photo–Fenton Activity in Mesoporous MIL–100(Fe) Decorated with ZnO Nanosphere for Pollutants Degradation. Appl. Catal. B Environ. 2019, 245, 428–438. [Google Scholar] [CrossRef]
  107. He, X.; Fang, H.; Gosztola, D.J.; Jiang, Z.; Jena, P.; Wang, W.N. Mechanistic Insight into Photocatalytic Pathways of MIL–100(Fe)/TiO2 Composites. ACS Appl. Mater. Interfaces 2019, 11, 12516–12524. [Google Scholar] [CrossRef] [PubMed]
  108. Oladipo, A.A. MIL–53 (Fe)–Based Photo–Sensitive Composite for Degradation of Organochlorinated Herbicide and Enhanced Reduction of Cr(VI). Process Saf. Environ. Prot. 2018, 116, 413–423. [Google Scholar] [CrossRef]
  109. Wang, Q.; Gao, Q.; Al–Enizi, A.M.; Nafady, A.; Ma, S. Recent Advances in MOF–Based Photocatalysis: Environmental Remediation under Visible Light. Inorg. Chem. Front. 2020, 7, 300–339. [Google Scholar] [CrossRef]
  110. Do, T.L.; Ho, T.M.T.; Doan, V.D.; Le, V.T.; Hoai Thuong, N. Iron–Doped Copper 1,4–Benzenedicarboxylate as Photo–Fenton Catalyst for Degradation of Methylene Blue. Toxicol. Environ. Chem. 2019, 101, 13–25. [Google Scholar] [CrossRef]
  111. Shi, S.; Han, X.; Liu, J.; Lan, X.; Feng, J.; Li, Y.; Zhang, W.; Wang, J. Photothermal–Boosted Effect of Binary Cu—Fe Bimetallic Magnetic MOF Heterojunction for High–Performance Photo–Fenton Degradation of Organic Pollutants. Sci. Total Environ. 2021, 795, 148883. [Google Scholar] [CrossRef]
  112. Zhong, Z.; Li, M.; Fu, J.; Wang, Y.; Muhammad, Y.; Li, S.; Wang, J.; Zhao, Z.; Zhao, Z. Construction of Cu–Bridged Cu2O/MIL(Fe/Cu) Catalyst with Enhanced Interfacial Contact for the Synergistic Photo–Fenton Degradation of Thiacloprid. Chem. Eng. J. 2020, 395, 125184. [Google Scholar] [CrossRef]
  113. Wu, Q.; Siddique, M.S.; Guo, Y.; Wu, M.; Yang, Y.; Yang, H. Low–Crystalline Bimetallic Metal–Organic Frameworks as an Excellent Platform for Photo–Fenton Degradation of Organic Contaminants: Intensified Synergism between Hetero–Metal Nodes. Appl. Catal. B Environ. 2021, 286, 119950. [Google Scholar] [CrossRef]
Figure 1. Classification of the different Fe–Cu based materials applied as catalysts in photo–Fenton processes.
Figure 1. Classification of the different Fe–Cu based materials applied as catalysts in photo–Fenton processes.
Catalysts 13 00159 g001
Figure 2. Reaction mechanism proposed for the degradation of pyridine by the UV/M–40%/H2O2 system. Reprinted with permission from Ref. [39]. Copyright 2020, Elsevier.
Figure 2. Reaction mechanism proposed for the degradation of pyridine by the UV/M–40%/H2O2 system. Reprinted with permission from Ref. [39]. Copyright 2020, Elsevier.
Catalysts 13 00159 g002
Figure 3. Schematic diagram illustrating the integration system of membrane filtration and the photo–Fenton process. BSA: bovine serum albumin; HA: humic acid. Reprinted with permission from Ref. [72]. Copyright 2019, Elsevier.
Figure 3. Schematic diagram illustrating the integration system of membrane filtration and the photo–Fenton process. BSA: bovine serum albumin; HA: humic acid. Reprinted with permission from Ref. [72]. Copyright 2019, Elsevier.
Catalysts 13 00159 g003
Figure 4. Schematic diagram for the photocatalytic degradation of MO by CZTS/ZFO p–n heterostructure plus H2O2: (a) the energy band structure, and the photo–induced carrier separation and transfer and (b) possible mechanism of photo–Fenton reaction in CZTS/ZFO + H2O2 photocatalytic system. Reprinted with permission from Ref. [84]. Copyright 2020, Elsevier.
Figure 4. Schematic diagram for the photocatalytic degradation of MO by CZTS/ZFO p–n heterostructure plus H2O2: (a) the energy band structure, and the photo–induced carrier separation and transfer and (b) possible mechanism of photo–Fenton reaction in CZTS/ZFO + H2O2 photocatalytic system. Reprinted with permission from Ref. [84]. Copyright 2020, Elsevier.
Catalysts 13 00159 g004
Figure 5. Possible mechanism of tetracycline degradation by CuFeO2/biochar in the heterogeneous photo–Fenton system. Reprinted with permission from Ref. [38]. Copyright 2020, Elsevier.
Figure 5. Possible mechanism of tetracycline degradation by CuFeO2/biochar in the heterogeneous photo–Fenton system. Reprinted with permission from Ref. [38]. Copyright 2020, Elsevier.
Catalysts 13 00159 g005
Figure 6. Possible mechanism of tetracycline degradation in the H–VL–PEF system. Reprinted with permission from Ref. [91]. Copyright 2020, Elsevier.
Figure 6. Possible mechanism of tetracycline degradation in the H–VL–PEF system. Reprinted with permission from Ref. [91]. Copyright 2020, Elsevier.
Catalysts 13 00159 g006
Figure 7. (A) Schematic illustration of the synthesis of FCCN. (B) Schematic diagram of the catalytic mechanism of FCCN (see text). Reprinted with permission from Ref. [91]. Copyright 2020, Elsevier.
Figure 7. (A) Schematic illustration of the synthesis of FCCN. (B) Schematic diagram of the catalytic mechanism of FCCN (see text). Reprinted with permission from Ref. [91]. Copyright 2020, Elsevier.
Catalysts 13 00159 g007
Figure 8. Schematic diagram of the synthesis process of MCuFe MOF. Reprinted with permission from Ref. [111]. Copyright 2020, Elsevier.
Figure 8. Schematic diagram of the synthesis process of MCuFe MOF. Reprinted with permission from Ref. [111]. Copyright 2020, Elsevier.
Catalysts 13 00159 g008
Figure 9. Photo–Fenton reaction mechanism of the charge transfer for hydroxyl radical generation over the MCuFe MOF under visible light irradiation. Reprinted with permission from Ref. [111]. Copyright 2020, Elsevier.
Figure 9. Photo–Fenton reaction mechanism of the charge transfer for hydroxyl radical generation over the MCuFe MOF under visible light irradiation. Reprinted with permission from Ref. [111]. Copyright 2020, Elsevier.
Catalysts 13 00159 g009
Figure 10. Proposed reaction mechanism for the photo–Fenton degradation of TCL over Cu2O/MIL(Fe/Cu). Reprinted with permission from Ref. [112]. Copyright 2020, Elsevier.
Figure 10. Proposed reaction mechanism for the photo–Fenton degradation of TCL over Cu2O/MIL(Fe/Cu). Reprinted with permission from Ref. [112]. Copyright 2020, Elsevier.
Catalysts 13 00159 g010
Table 1. Summary of recently used Fe–Cu ferrites and mixed ferrites as photo–Fenton photocatalysts in wastewater treatment.
Table 1. Summary of recently used Fe–Cu ferrites and mixed ferrites as photo–Fenton photocatalysts in wastewater treatment.
Material (a)ConditionsContaminantDegradation EfficiencyReference
Fe2O3/CuFe2O4(b)UV light, pH 7,
[H2O2] = 832.50 mg L−1
Pyridine (100 mg L−1)>99% within 30 min; TOC removal of 97% within 50 min[39]
α–Fe2O3/CuFe2O4 (c)Natural solar light, pH 7, [H2O2] = 300 mg L−1MB (MB, 100 mg L−1)100% removal of the dye in 180 min[62]
CuFe2O4/CuO (d)Halogen lamp, pH not indicated, [H2O2] not indicatedMB (40 mg L−1), Rhodamine B (20 mg L−1), Methyl Orange (20 mg L−1)100% dye degradation within 50 min[63]
Hollow CuFe2O4 nanospheres (e)300–W UV curing lamp (λ > 400 nm), pH not indicated, [H2O2] = 0.02 MMB (30 mg L−1)96.4% in 60 min[40]
CuFe2O4/biochar nanocomposites (f)Fluorescent lamp (395−580 nm)/Sunlight, pH 3–7, [H2O2] from 2.5 to 10 μMRhodamine B100% color removal was obtained at 10 and 20 min of reaction for 10 and 50 mg L−1 of dye with solar radiation[64]
Bi2WO6/CuFe2O4 (e)Visible light, pH 2.6–6.3, [H2O2] = 10 mMTetracycline hydrochloride (TCH)After 30 min, 92.1% TCH (20 mg L−1) degradation (pH 2.6) efficiency and 50.7% and 35.1% mineralization performance.[65]
CuFe2O4/tartaric acid (TA) (b)UV–curing lamp (365–450 nm), pH 5.0, [H2O2] = 0.02 MMB (50 mg L−1)Introducing TA enhanced MB decolorization rate from 52.0% to
92.1% within 80 min.
[67]
CuFe2O4 (f)Sunlight, pH 7.0, [H2O2] = 0.3 MMB (10 mg L−1)Decolorization efficiency was 96.1% in 45 min of reaction.[69]
CuFe2O4@rGO (g)Simulated sunlight, pH not indicatedGuaiacylglycerol–β–guaiacyl ether (lignin model compound)Yields of 72.6% and 52.5% were achieved for guaiacol and 2–methoxy–4–propylpheno, respectively, in 60 min.[70]
CuFe2O4 nanoparticles doped in polyvinylidene fluoride (PVDF) membranes (h)Xe arc lamp, pH 3.0–11.0. The H2O2 (30 wt.%) dosage was in the range 50–1200 μL in 50 mL MB solutionMB (100 mg L−1)MB was thoroughly degraded in
30 min when the pH is 3.0, while there is still 15.6% of the MB left at solution pH of 11.0.
[72]
Cu0.5Mn0.5Fe2O4 (i)Cu0.5Mn0.5Fe2O4 0.08 g L−1
pH: 4.2
[H2O2] = 10 mM
BPA (10 mg L−1)100% degradation and 47.6% mineralization
5 min
[83]
Cu0.8Mn0.2Fe2O4 (e)Catalyst 0.100 g L−1
pH 3 and 11
300 W Xe lamp with a 420 nm UV–cut off filter
TC–HCL
100 mL
80 mg L−1
99%
30 min
[82]
CZTS/ZFO
p–n heterostructures (e)
Catalyst 0.5 g L−1
pH 3 to 9
[H2O2] = 10 mM.
500 W Xe high intensity discharge lamp > 450 nm
MO
10 mg L−1
91%
120 min
At optimum pH: 6
[84]
CuFeMnO4 on the surface of a honeycomb ceramic substrate (j)0.05g of catalyst
pH = 6.71.
[H2O2] from 0 to 0.1 M
Solar light
Phenol as a VOC model 10, 20,50 and 100 mg L−1
MB 10 mg L−1
∼99.18% COD and MB was removed
20 min
[85]
(a) The synthesis method is shown in the footnotes. (b) Sol–gel method. (c) Modified Pechini method. (d) Solution combustion synthesis method. (e) Solvothermal method. (f) Co–precipitation. (g) Solvent–assisted interfacial reaction process. (h) Non–solvent induced phase separation method. (i) Chemical co–precipitation method followed by high–temperature annealing treatment. (j) See original papers for detail.
Table 2. Summary of recently used CuFeO2 and CuFeS2 as photo–Fenton photocatalysts in wastewater treatment.
Table 2. Summary of recently used CuFeO2 and CuFeS2 as photo–Fenton photocatalysts in wastewater treatment.
Material (a)ConditionsContaminantDegradation EfficiencyReference
Delafossite–type CuFeO2 (b,c)visible light, pH from 2.4 to 3.6, H2O2 from 3 to 13 mM, catalyst from 0.13 to 0.33 gL−1Reactive Red 141 dye, 50 mg L−1about 98% at 30 min[86]
3R–delafossite CuFeO2 microcrystals (b)200 mL reactor, 20 mM of H2O2, 1 g L−1 catalyst, initial pH 8Tetracycline hydrochloride 20 mg L−196.1% in 180 min[87]
CuFeS2(c)visible light, 20 mM of H2O2, 0.2 g L−1 catalyst, pH 6bisphenol A (BPA) 20 mg L−197% in 60 min[88]
CuFeS2 chalcogenide powders (b,c)visible light, pH 3.0, 8.33 mM of H2O2, 0.2 g L−1 of catalysttartrazine, 100 mg L−199.1% of tartrazine decolorization after 40 min and 87.3% of mineralization after 150 min[89]
MAMR–Fe–Cu@SiO2/SC, a core–shell structure of CuFeO2@SiO2/starch–derived carbon anchored on a Manganese Residue (d)Xe–lamp with UV cut–off, pH from 2.5 to 9.5, 15 mM of H2O2, 0.7 g L−1 of catalystTetracycline 50 mg L−1100% in 40 min[90]
CuFeO2/biochar (b)Xe Lamp with UV cutoff, pH 4 to 8, 20 mM of H2O2, 0.2 g L−1 of catalystTetracycline 20 mg L−197.6% in 120 min[38]
CuFeO2/biochar (b)photo–electro–Fenton, Xe–Lamp with UV cutoff, pH from 3 to 11, H2O2 generated by a NO–doped/porous carbon cathodeTetracycline (20 to 200 mg L−1)100% in 60–70 min[91]
Nitrogen/oxygen self–doped porous biochar (NO/PBC) and NO/PBC–supported CuFeO2 (CuFeO2–NO/PBC) (b)undivided quartz reactor, visible light, pH from 3 to 11, H2O2 by a gas diffusion electrodeTetracycline 20 mg L−198% at 30 min[92]
(a) The synthesis method is shown in the footnotes. (b) Hydrothermal method. (c) Microwave–assisted hydrothermal method. (d) Reductive roasting and mechanical activation.
Table 3. Summary of recently used Fe–Cu oxides composites as photo–Fenton photocatalysts in wastewater treatment.
Table 3. Summary of recently used Fe–Cu oxides composites as photo–Fenton photocatalysts in wastewater treatment.
Material (a)ConditionsContaminantDegradation EfficiencyReference
Iron–copper oxide impregnated NaOH–activated biochar
(FeCu/ABC catalyst) (b)
Heterogeneous PEF process,
pH = 5.8
catalyst dosage of 1 g L−1, electrical current of 200 mA
CIP (45 mg L−1)100% removal
2 h
[95]
Fe20Cu80(0.2 wt.%)/D3 (c)catalyst ~200 mg L−1,
[H2O2] (200 mg L−1; 5.88 mM), 20 °C,
simulated sunlight. pH = 4
Phenol
(100 mg L−1)
90% removal
2 h
[103]
NP–3 (CuII0.4FeII0.6FeIII2O4) (d)NP–3 = 400 mg L−1, [H2O2] = 1.76 × 10−1 mol L−1, pH = 7.5
Optonica SP1275 LED lamp (GU10, 7 W, 400 Lm, 6000 K, Optonica LED, Sofia, Bulgaria)
MB
(1.5 × 10−5 mol L−1)
RhB
(1.75 × 10−5 mol L−1)
100%
140 min
[98]
Iron oxide (α–Fe2O3, hematite) colloids combined with other transition–metal oxide (TMO) colloids (e.g., CuO and ZnO) (e)750 mg L−1 catalyst, [H2O2] = 0.025 mol L−1. Tungsten halogen lampsMO
(25 µM)
7 to 78%
60 min
[99]
Fe–Mn–Cu@ATP (f)500 mL of Fe–Mn–Cu@ATP dosage (1–12 g L−1),
pH = 3
[H2O2] (0.1–0.6 mol L−1)
UV 40 W UV lamp
pharmaceutical wastewaterCOD removal: 64.9%
180 min
[100]
Fe2O3–CuO@Ca–Alg hydrogel (g)4.7 mg L−1 of catalyst,
pH = 3.5
[H2O2] = 33.17 mmol L−1
UV light
[MNZ]0
10 mg L−1
Removal = 95%
85 min
[101]
FCCN (h)Catalyst 1 g L−1
pH: 7–12
[H2O2] = 15 mM
500W Xe lamp
564 nm cut–off filter
MO, acid orange II and mordant yellow 10–20 mg L−1100%
15 min
[102]
(a) The synthesis method is shown in the footnotes. (b) Pyrolysis of activated biochar followed by impregnation. (c) Deposition of the NPs (Fe, Cu, or Fe–Cu) onto the surface of commercial carbonaceous supports (d) Co–precipitation. (e) Hydrothermal method. (f) Aeration–coprecipitation method. (g) Green method using walnut green shells. (h) See Figure 7A.
Table 4. Summary of recently used Fe–Cu bimetallic MOFs as photo–Fenton photocatalysts in wastewater treatment.
Table 4. Summary of recently used Fe–Cu bimetallic MOFs as photo–Fenton photocatalysts in wastewater treatment.
Material (a)ConditionsContaminantDegradation EfficiencyReference
Fe–CuBDC(b)Simulated sunlight, pH 6,
[H2O2] = 50 mM,
[Catalyst] = 1 g L−1
Methyl Blue
(50 mg L−1)
100% removal of the dye in 70 min[110]
MCuFe MOF (b,c)300 W xenon lamp equipped with a UV cut–off filter (λ > 400 nm), pH 4–9, [H2O2] = 5 mM,
[Catalyst] = 0.05 g L−1
Methyl Blue
(50 mg L−1)
100% removal of the dye in 40 min[111]
Cu2O/MIL(Fe/Cu) (b)500 W xenon lamp, pH 7.47, [H2O2] = 49 mM,
[Catalyst] = 0.5 g L−1
Thiacloprid
(80 mg L−1)
100% removal of TCL in 20 min; 82% TOC removal in 80 min[112]
L–MIL–53 (Fe, Cu) (b)300 W xenon lamp equipped with a UV cut–off filter (λ > 420 nm), pH 7, [H2O2] = 5 mM,
[Catalyst] = 0.1 g L−1
Ciprofloxacin
(20 mg L−1)
60% removal of CIP in 30 min[113]
(a) The synthesis method is shown in the footnotes. (b) solvothermal method. (c) Hydrothermal method.
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

Bosio, G.N.; García Einschlag, F.S.; Carlos, L.; Mártire, D.O. Recent Advances in the Development of Novel Iron–Copper Bimetallic Photo Fenton Catalysts. Catalysts 2023, 13, 159. https://doi.org/10.3390/catal13010159

AMA Style

Bosio GN, García Einschlag FS, Carlos L, Mártire DO. Recent Advances in the Development of Novel Iron–Copper Bimetallic Photo Fenton Catalysts. Catalysts. 2023; 13(1):159. https://doi.org/10.3390/catal13010159

Chicago/Turabian Style

Bosio, Gabriela N., Fernando S. García Einschlag, Luciano Carlos, and Daniel O. Mártire. 2023. "Recent Advances in the Development of Novel Iron–Copper Bimetallic Photo Fenton Catalysts" Catalysts 13, no. 1: 159. https://doi.org/10.3390/catal13010159

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