Next Article in Journal
Bioenergy Generation and Wastewater Purification with Li0.95Ta0.76Nb0.19Mg0.15O3 as New Air-Photocathode for MFCs
Next Article in Special Issue
Photocatalytic Degradation of Methyl Orange Dyes Using Green Synthesized MoS2/Co3O4 Nanohybrids
Previous Article in Journal
Morphology Effects on Structure-Activity Relationship of Pd/Y-ZrO2 Catalysts for Methane Oxidation
Previous Article in Special Issue
Photocatalytic Activity of ZnO/Ag Nanoparticles Fabricated by a Spray Pyrolysis Method with Different O2:N2 Carrier Gas Ratios and Ag Contents
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Magnetic TiO2/Fe3O4-Chitosan Beads: A Highly Efficient and Reusable Catalyst for Photo-Electro-Fenton Process

1
BIOSUV Research Group, INTECX Building, University of Vigo, Lagoas-Marcosende Campus, 36310 Vigo, Spain
2
Research Unit in Electrochemistry, Materials and Environment (UR16ES02), IPEIK, University of Kairouan, Kairouan 3100, Tunisia
*
Author to whom correspondence should be addressed.
Catalysts 2022, 12(11), 1425; https://doi.org/10.3390/catal12111425
Submission received: 16 October 2022 / Revised: 7 November 2022 / Accepted: 9 November 2022 / Published: 13 November 2022
(This article belongs to the Special Issue Structured Semiconductors in Photocatalysis)

Abstract

:
Heterogeneous photo-electro-Fenton process is an attractive technology for the removal of recalcitrant pollutants. To better exploit the presence of an irradiation source, a bifunctional catalyst with TiO2 nanoparticles embedded into an iron–chitosan matrix was developed. The catalytic activity of the catalyst was improved by the optimization of the loaded TiO2 content. The prepared composite catalysts based on TiO2, Fe3O4 and chitosan were called TiO2/Fe3O4-CS beads. The best catalyst with an optimal ratio TiO2/Fe = 2 exhibited a high efficiency in the degradation and mineralization of chlordimeform (CDM) insecticide. Under the optimum conditions (concentration of catalyst equal to 1 g L−1 and applied current intensity equal to 70 mA), a real effluent doped with 30 mg L−1 of CDM was efficiently treated, leading to 80.8 ± 1.9% TOC reduction after 6 h of treatment, with total removal of CDM after only 1 h.The generated carboxylic acids and minerals were identified and quantified. Furthermore, the stability and reusability of the developed catalyst was examined, and an insignificant reduction in catalytic activity was noticed for four consecutive cycles of the photo-electro-Fenton process. Analyses using SEM, XRD and VSM showed a good stability of the physicochemical properties of the catalyst after use.

1. Introduction

In recent decades, the extensive use of pesticides to improve agricultural production has led to an increased risk of water pollution. In fact, these pollutants can be classified as persistent and extremely toxic organic substances. The ability of these harmful pollutants to easily bioaccumulate even at very low concentrations represents a serious issue that can cause problems for the environment, human health and living organisms [1]. Consequently, the removal of pesticides from water and wastewater has become a great global challenge.
In recent years, electrochemical advanced oxidation processes (EAOPs) have been considered an option for the removal of toxic and persistent organic micropollutants from wastewater [2,3,4]. One of the most powerful and attractive techniques among EAOPs is the electro-Fenton (EF) process, whereby hydrogen peroxide (H2O2) is generated at a cathode by O2 reduction (Equation (1)), and a ferrous ion or iron oxide catalyst is added to the effluent [5]. The EF process is very effective in the mineralization of organic pollutants due to the production of strongly oxidizing hydroxyl radicals through the Fenton reaction (Equation (2)) [6]. However, this technology is limited by a few factors, such as the rapid accumulation of Fe(III) ions and the possibility of their complexation with hydroxyl ions and oxidation products during treatment [7].
O 2 + 2   H + + 2   é     H 2 O 2
H 2 O 2 +   Fe 2 +     Fe 3 + +   OH + OH
The presence of UV irradiation has two specific aims: the regeneration of Fe(II) ions and the degradation of the formed complexes until their mineralization due to an additional production of •OH [8]. This process is called the photo-electro-Fenton (PEF) process, and it leads to great performance in the mineralization of organic pollutants [9,10].
However, the application of UV irradiation could increase the operational cost, but it can be justified by the use of a photocatalyst, such as TiO2, to increase its economic cost effectiveness. The application of a ferromagnetic-TiO2 composite material is widely used for photocatalysis due to its efficiency and magnetic properties [11,12].
In fact, in a heterogeneous system, many solid catalysts containing iron, such as Fe, Fe2O3, Fe3O4 and FeOOH, were successfully used for the treatment of recalcitrant organic pollutants in water [13,14,15,16,17]. Among iron oxides, the immense popularity of Fe3O4 as a catalyst originates from its broad application potential due to high saturation magnetization, easy handling, relatively low cost, non-toxicity and environmentally friendly character [18,19]. Titanium oxide (TiO2) has also shown widespread photocatalytic application in the field of wastewater treatment due to its unrivalled properties of non-toxicity, easy UV activation, chemical stability and availability [20]. Despite all these outstanding properties, TiO2 deployment for photocatalytic application has witnessed drawbacks due to its large energy band gap of about 3.2 eV [20,21]. In addition, one of the problems hindering TiO2 use is its agglomeration in aqueous solution, which requires a post-filtration treatment [21]. In order to overcome the problem of the band gap of TiO2, several studies revealed that the presence of Fe3O4 enhances the photocatalytic activity of TiO2 by decreasing the charge carrier recombination, with a band gap of the composite material around 2.5 eV. This is caused by the generated Z-scheme on the composite TiO2/Fe3O4 where the recombination between holes and electrons on TiO2 is restrained by the addition of another valence band (Fe3O4). Indeed, Fe-species have been reported as an outstanding alternative for creating Z-scheme photocatalytic heterojunction systems [22]. Moreover, this modified TiO2 nanoparticle can be separated from water by means of an external magnetic field [11,23]. However, the use of magnetic separation is complicated in real water applications. Therefore, the use of a catalytic support, such as chitosan, for both metal oxides (Fe3O4 and TiO2) not only avoids the agglomeration issue, but it also facilitates catalyst recovery, protects ferroparticles from oxidation and extends their storage life [24].
Therefore, in the present study, we innovatively synthesized the TiO2/Fe3O4-CS catalytic beads through a green co-precipitation method in only one step. The co-precipitation method involves the dispersal of a mixture of chitosan and metals into an alkaline solution to form nanoparticles strongly bound to chitosan with attractive catalytic and magnetic properties for easy catalyst recovery and reuse [24,25,26,27,28]. TiO2/Fe3O4-CS was used as a novel nano-structured heterogeneous catalyst for the oxidation of CDM insecticide by an efficient cyclic PEF process; therefore, Fe(II), within the structure of Fe3O4, can activate the Fenton system [29], and TiO2 acts as a photocatalyst to better exploit the use of UV irradiation. In the PEF system, an electrochemical reactor was used for the electrogeneration of H2O2, and the catalyst was added to a second photocatalytic reactor. The effects of operational parameters were studied, such as the TiO2 loading in the beads, catalyst dosage and the current intensity applied. Under optimal conditions, aqueous wastewater doped with 30 mg L−1 of CDM was treated. The carboxylic acids and inorganic ions generated during the treatment process were identified and quantified. Finally, the electric energy consumed was estimated, and the stability of the best catalyst in terms of CDM and TOC removal efficiencies was studied.

2. Results

2.1. Optimization of Operational Parameters

2.1.1. Performance Comparison of Several Processes for the Removal of CDM

Initially, some control experiments were performed. To begin with, the adsorption assays of the developed catalytic beads showed no CDM adsorption after 4 h of treatment on the synthesized catalytic beads (data not shown). Then, the low efficiencies of photoelectrolysis–H2O2 (electrogenerated H2O2 + UV-LED) and photocatalysis (1 g of TiO2/Fe3O4-CS catalyst + UV-LED) systems revealed, respectively, 70.6 ± 5.3% and 13.2 ± 4.4% of CDM removal yield after 1 h of electrolysis (Figure 1) along with only 18.2 ± 3.3% and 4.4 ± 1.6% of TOC abatement after 4 h of treatment (Table 1). However, the PEF process running at I = 50 mA and pHi = 3, in the presence of 1 g L−1 TiO2/Fe3O4-CS, exhibited higher efficiency, with total degradation of CDM after only 1 h and TOC abatement equal to 69.6 ± 2.7% after 4 h of treatment (Table 1).

2.1.2. Influence of TiO2 in Magnetic Chitosan Beads

As found previously in the PEF system, the composite catalyst TiO2/Fe3O4-CS showed higher catalytic activity, with total degradation of CDM after only 1 h and TOC abatement of around 70% after 4 h of treatment. However, in the presence of Fe3O4-CS catalytic beads, only 87.9 ± 3.1% CDM removal yield and TOC abatement was achieved, not exceeding 33.9 ± 4.1%. The higher photocatalytic activity of TiO2/Fe3O4-CS beads could be attributed to the acceleration of electron mobility in the FeIII/II/TiO2 system [12,30] and the synergistic combination of PEF with photocatalysis. It is widely reported that the combination of TiO2 and Fe3O4, with their different band gaps, can suppress the electron–hole recombination, enhancing the photocatalytic activity, and thus, the PEF process efficiency, thanks to the generation of the Z-scheme [11,22,30,31,32]. Moreover, the degradation of CDM by the PEF process was found to follow a pseudo-first-order kinetic (Figure 1b), which is probably related to the steady •OH concentration throughout the treatment process [33,34]. To evaluate the catalytic decomposition of the oxidant, the concentration of H2O2 was monitored during treatment with and without the developed catalysts, and the results are presented in Figure 1c. In the absence of the catalyst and UV-LED lamp, the accumulated amount of H2O2 after 4 h of electrolysis was estimated at 64.4 ± 0.5 mg L−1. In the presence of UV-LED irradiation, the concentration of H2O2 decreased to 53.8 ± 0.9 mg L−1, which means that the amount of H2O2 decomposed by photolysis after 4 h of electrolysis was around 16.6%. As expected, by adding the Fe3O4-CS and TiO2/Fe3O4-CS catalytic beads to the (H2O2 + UV-LED) system, the concentration of the oxidant decreased significantly to 39.9 ± 1.1 mg L−1 and 29 ± 0.8 mg L−1, respectively, indicating that approximately 38% and 55% of the H2O2 amount was decomposed. These findings were in agreement with our previous results, and they confirmed the best catalytic activity of TiO2/Fe3O4-CS beads because an increase in the decomposition rate of H2O2 is related to an increase in the production yield of •OH radicals [35]. Therefore, it is clear that the incorporation of TiO2 into magnetic CS beads favors an effective usage of H2O2, allowing great improvement to the PEF process efficiency.

2.1.3. Effect of TiO2 Loading Content into Chitosan Beads

In order to increase the catalytic activity of TiO2/Fe3O4-CS beads, the effect of TiO2 content in the catalyst was studied. For this purpose, the molar ratio TiO2/Fe was varied during the preparation of the catalyst from 1 to 3 (to facilitate a designation of the developed catalysts, the molar ratio values between the brackets were added to TiO2/Fe3O4-CS). The results showed that a molar ratio equal to 2 (TiO2/Fe3O4-CS(2)) was optimal, leading to total degradation of CDM after 1 h (Figure 1d) and TOC removal yield of around 70% after 4 h of treatment (Table 1), against 38.2 ± 3.5% and 54.2 ± 5.6%, respectively, for the molar ratios TiO2/Fe = 1 and 3. On the other hand, the comparison of the photocatalytic activities of TiO2/Fe3O4-CS(2) and TiO2-CS (Figure 1d) showed a remarkable difference in their catalytic performance, which was greater for the composite catalyst. In fact, the CDM degradation rate did not exceed 35% after 1 h of treatment using TiO2-CS beads, with low TOC abatement estimated at 24.5 ± 1.9% after 4 h of treatment. Therefore, it is obvious that the Fe(II) ions in TiO2/Fe3O4-CS catalytic beads notably improve their catalytic performance by the Fenton reaction with H2O2, so that •OH radicals are produced. Numerous studies have reported the high photocatalytic activity of the mixture of TiO2/Fe3O4 compared to pure TiO2, and they explained this improvement by the fast photogenerated electron transfer between Fe3O4 and TiO2, which can effectively reduce electron/hole recombination [12,33,36,37].
To sum up, a molar ratio TiO2/Fe = 2 is optimal both for the removal of the pollutant and its mineralization. In fact, varying the molar ratio from 1 to 2, the photocatalytic activity of the catalytic beads increased through the increase in TiO2 active sites. The improvement in pollutant degradation and mineralization by the increase in TiO2 load in the iron catalysts has already been reported by many research studies, and it is related to the enhancement of the electron transfer reducing electron–hole recombination [32,38]. For a molar ratio TiO2/Fe = 3, the catalytic performance of the system decreased due to the decrease in active iron sites in favor of the photocatalytic sites of TiO2, which has a wide band gap energy, causing a rapid recombination of the electron–hole pairs [39]. It is well known that the fast recombination of charge carriers significantly lowers the photocatalytic performance [24,40]. Thus, a TiO2/Fe molar ratio equal to 2 was optimal for the preparation of TiO2/Fe3O4-CS beads, and the catalyst TiO2/Fe3O4-CS(2) was selected thereafter.
The photodegradation of pesticides using TiO2/Fe3O4-based materials as catalysts has been poorly reported. Compared to the efficiency of TiO2/Fe3O4 composite catalysts listed in Table 2, TiO2/Fe3O4-CS showed good catalytic activity. Nevertheless, the reactor set-up, such as the irradiation source, can greatly affect the overall efficiency of the process [41].

2.1.4. Effect of Catalyst Dosage

Catalyst dosage was another parameter affecting the heterogeneous Fenton reaction [45]. The effect of catalyst dosage on the degradation of CDM by the PEF process was studied by applying a current intensity of 50 mA at pHinitial = 3. It is clear from Figure 2a that the increase in TiO2/Fe3O4-CS(2) catalytic beads from 0.25 g L−1 to 1 g L−1 increased the rate of CDM removal, and an almost total degradation was obtained for all assays after 1 h of treatment, with an improvement in TOC abatement, respectively, from 40.9 ± 4.1% to 69.6 ± 2.7% (Table 1). For a concentration equal to 1.5 g L−1, there was a trivial enhancement in CDM and TOC removal efficiencies. However, the removal yields of CDM and TOC were decreased for a high concentration equal to 2 g L−1. Indeed, an increase in the concentration of catalytic beads in solution can inhibit the penetration of light into the photocatalytic reactor [46,47]. Additionally, an increase in the active catalytic sites can cause secondary reactions with •OH radicals, thus producing less reactive oxidizing species (Equation (3)), which can reduce the efficiency of the process for the mineralization of organic pollutants [25].
Fe 2 + + OH     Fe 3 + +   OH

2.1.5. Effect of the Current Intensity

The current intensity applied is a key parameter in the electrochemical Fenton technologies because it is the driving force for the reduction in oxygen, leading to the generation of H2O2 at the cathode, and it affects the regeneration of Fe(II) [48]. In order to study the effect of the current intensity applied on CDM degradation and its mineralization by the PEF process, several experiments were carried out using different current intensities in the range from 50 mA to 100 mA in the presence of 1 g of TiO2/Fe3O4-CS(2) and at initial acid pH (Figure 2b). The results showed that an increase in the current intensity from 50 mA to 70 mA leads to total degradation of the pollutant after only 1 h, with almost 10% improvement in TOC abatement (Table 1). This finding can be mainly attributed to the fast production of H2O2 at a higher current and the fast regeneration of Fe(II) enhancing the production of •OH radicals [49].
As seen in Figure 2b and Table 1, no further increase in the removal efficiencies was observed for the current intensity applied beyond 70 mA. This behavior can indicate that parasitic reactions, such as the four-electron reduction in O2 with H2O formation, as well as the decomposition and hydrogenation of H2O2, would take place when the current increased beyond a certain value [49,50,51,52]. On the other hand, working at higher current intensities can generate a degradation of our organic catalytic support “CS”. In fact, several studies have shown the degradation of CS in the presence of an irradiation source and high concentrations of H2O2 [53,54]. An optimal current intensity equal to 70 mA, which ensured high treatment efficiency and good catalyst stability, was therefore established for the upcoming experiments.
To confirm the role of •OH radicals in the mineralization of the organic pollutant, isopropanol was used as a •OH scavenger. The degradation of CDM was totally inhibited in the presence of isopropanol. Thus, •OH should be the main active species for the degradation of CDM by the PEF process.

2.2. Treatment of Wastewater Doped with Chlordimeform by Photo-Electro-Fenton Process

To evaluate the applicability of the PEF process using TiO2/Fe3O4-CS(2) as a catalyst for the removal of CDM in a more complex matrix than ultrapure water, an experiment was carried out on secondary treated wastewater kindly given by the municipal wastewater treatment plant in the northwest of Spain, whose physiochemical characteristics are summarized in Table 3.
Under optimal experimental conditions ([CDM]i = 30 mg L−1, [TiO2/Fe3O4-CS(2)] = 1 g L−1, and I = 70 mA), total degradation of CDM was achieved after 1 h, and TOC abatement reached 50.6 ± 5.1% after 1 h of electrolysis; then, it increased slowly to attain 80.8 ± 1.9% after 6 h of treatment (Figure 3b).
Furthermore, a slight decrease in the efficiency of the system was noticed compared to previous results when ultrapure water was used as the matrix (Figure 2b). This finding can be explained, according to zazouli et al. [55], by the presence of organic matter and inorganic anions in wastewater. In fact, the anions (X) can react with the •OH to produce X−• radicals, which are less reactive.
In addition to the aqueous matrix influence, it should be noted that EEC is another big concern, which should be considered for sustainable development [56]. The results showed that global EEC was low compared to some literature values on the treatment of real wastewater by electrochemical processes [57,58,59] around 1.7 kWh m−3 and 1.4 kWh m−3, respectively, for real wastewater and ultrapure water matrices. However, it should be noted that the difference (≃0.3 kWh m−3) is probably due to the presence of degradable compounds in the treated wastewater matrix, whose oxidation requires a little more energy.

2.3. Evaluation of Organic Acids and Minerals Produced during Treatment

Studies of organic pollutants’ mineralization by advanced oxidation processes show that oxidation by •OH radicals leads to the formation of carboxylic acids [60,61]. To better assess the oxidative capacity of our PEF process, an analysis of carboxylic acids was performed in order to identify and quantify them during treatment. The evolution of the identified acids during the mineralization of CDM is presented in Figure 3c. All the acids produced reached their maximum concentrations after around 2 h. Then, they decreased to a zero value after 6 h of treatment, indicating the deep mineralization of all detected acids. The concentration of oxalic acid during the treatment process was low (almost zero), which can be explained by the high reactivity of oxalic acid with Fe(III) and the formation of Fe(III)–oxalate complexes that are easily photolysed in the presence of UV light and H2O2 oxidant [62]. However, the other acids detected were malonic, succinic, formic and glycolic acids, appearing at the beginning of treatment. These results showed the effectiveness of our PEF process in removing the aliphatic compounds known for their resistance to oxidation.
On the other hand, considering the CDM molecule is composed of a chlorine atom and two nitrogen atoms, its treatment by an advanced oxidation process leads to the production of minerals. The evolution of these ions was monitored by ion chromatography. The obtained results (Figure 3d) showed that the chloride ions reached a maximum concentration of 6.22 mg L−1 after 1 h; this concentration presented the total amount initially present in the parent molecule. This finding was in agreement with several studies, which have shown that the release of chloride ions is rapid using a BDD anode and that the dechlorination of aromatic compounds occurs before the opening reaction of the aromatic cycle [63,64,65]. Then, the concentration of chloride ions decreased to approximately half (3.05 mg L−1) after 6 h of treatment, which was in agreement with the results found by Mhemdi et al. [64] who studied the effect of the anode material (platinum and BDD) on the evolution of chlorides during the mineralization of 2-chlorobenzoic acid by an electrochemical oxidation process. The results showed that, for platinum, the concentration of chloride ions reached a maximum after 2 h and then remained constant. Using the BDD anode, a different behavior of chloride ions was observed. They accumulated rapidly after 1 h; then, their concentration decreased due to their oxidation to chlorine (Cl2), which transformed into hypochlorite ions HClO by hydrolysis. Consequently, the decrease in the accumulated chloride concentration during treatment could be explained by the oxidizing power of the BDD anode and the synthesized catalysts, causing the transformation of Cl ions into Cl2.
Furthermore, the concentration of nitrate ions (NO3) after 6 h was estimated at 0.68 mg L−1. However, the accumulated concentration of ammonium ions (NH4+) was high, and it was around 4.75 mg L−1. The accumulated concentrations of nitrate and ammonium ions after 6 h of treatment represent approximately 90% of the amount of nitrogen initially contained in the molecule of CDM. Consequently, one can confirm that CDM is mineralized during PEF application.

2.4. Stability of the Catalyst

2.4.1. Catalytic Stability

The recycling and reusability of a heterogeneous catalyst are important parameters for economic and environmental considerations [66]. The TOC abatement obtained using the same TiO2/Fe3O4-CS(2) beads for four consecutive runs is depicted in Figure 4. After each run, the catalyst was washed with distilled water and dried at ambient temperature. The results showed that TOC removal remained almost constant for four cycles of reuse. The slight decrease (≃6.7%) observed in the fourth cycle could be attributed to the modification of the physicochemical properties of the catalyst’s surface by the effect of UV-LED irradiation and the presence of H2O2 as an oxidant. In addition, possible leaching of Fe ions in the reaction medium was checked. The leached amount of Fe after each run did not exceed 1.7 mg L−1, corresponding to 0.5% of the initial metal content in catalytic beads. These results revealed an excellent structural stability of the catalytic beads and minimal leaching of iron according to the environmental standards demanded by the European Union (<2 mg L−1) [67].

2.4.2. Characterization of Catalyst before and after Use

SEM Analysis

The SEM images of TiO2/Fe3O4-CS(2) beads before and after use are illustrated in Figure 5. As shown, the surface of the beads is smooth, indicating that iron and TiO2 are attached to CS. The use of high magnification highlighted the observation of TiO2 nanoparticles on the surface of the beads. The comparison of images in Figure 5a,b showed that, after four consecutive cycles, the surface of the beads was slightly damaged by oxidizing conditions and UV-LED irradiation, which explained the leaching of the metal.

XRD Analysis

In order to determine the crystalline structure of the TiO2/Fe3O4-CS(2) beads, an XRD analysis was performed. In the XRD patterns (Figure 6a), the peak positions at 2Theta = 30.14°, 35.59°, 43.47°, 53.87°, 57.41° and 62.65° corresponded, respectively, to the Miller indices of (220), (311), (400), (422), (511) and (440) of the crystalline structure of Fe3O4, and the peak positions at 2Theta = 25.29°, 37.78°, 38.59°, 48.01°, 55.03°, 68.74°, 70,25°, 75.01°, 82.62° could be attributed to the diffraction planes of (110), (101), (004), (220), (105), (400), (220), (215) and (312) of the crystalline structure of TiO2. The obtained peaks are well matched with standard data JCPDS-ICDD card N°:00-019-0629 for the Fe3O4 magnetite and 00-021-1272 for the TiO2 anatase.
The analysis of the catalytic beads used allowed us to obtain a diffractogram similar to that of fresh beads. In fact, indexing the pattern exhibited the crystal structure of the TiO2 anatase (ICDD 00-021-1272). However, the peak positions at 2Theta = 30.14°, 35.59°, 43.47°, 48.01°, 55.03°, 57.41° and 62.65° corresponded, respectively, to the Miller indices of (220), (311), (400), (421), (422) (511) and (440) of the crystalline structure of maghemite Fe2O3 (ICDD 00-039-1346) instead of magnetite Fe3O4. Fe2O3 had the same spinel structure as magnetite, but it only contained iron in the trivalent Fe(III) state [68], which indicated the oxidation of Fe3O4. Furthermore, using the Debye–Scherrer equation, the grain size of TiO2 was estimated at 64.5 nm, while the size of Fe3O4 was around 11.6 nm, indicating that the beads exhibit superparamagnetic behavior [69].

Magnetic Properties Analysis

The magnetization curve of the TiO2/Fe3O4-CS(2) beads revealed a saturation intensity equal to 11.40 emu/g (Figure 6b). This low value is due to the presence of non-magnetic materials (CS and TiO2) [70]. The magnetization intensity of the beads used was almost the same, and it was estimated at 10.48 emu g−1, which means that the beads have a good stability of the magnetic properties after four consecutive cycles.

3. Materials and Methods

3.1. Chemical Products

All chemicals were of analytical–laboratory grade and applied without further purification. Chlordimeform, chitosan, sodium hydroxide, ferrous sulfate iron (II), iron (III) chloride, titanium oxide and potassium titanium oxide oxalate dehydrate were purchased from Sigma-Aldrich (Madrid, Spain). Acetic acid, sulfuric acid and nitric acid were supplied by Analar Normapur (Radnor, PA, USA). Acetonitrile (HPLC-grade) was purchased from Fisher Scientific (Loughborough, UK). Ultrapure water obtained through reverse osmosis technology (Basic 360) was utilized throughout all the experiments.

3.2. Preparation of TiO2/Fe3O4-Chitosan Magnetic Beads

An eco-friendly, low-cost and simple approach was used to synthesize the composite catalysts. Magnetic TiO2/Fe3O4-CS beads were prepared via a precipitation method using sodium hydroxide (1 M) as a precipitating agent. First, 2% CS gel solution was prepared by dissolving 1 g of CS in acetic acid (1%) under stirring at room temperature. After the total dissolution of CS flakes, 5 mmol of iron salts at a molar ratio Fe3 +:Fe2 + = 2:1 was added. Then, the TiO2 nanoparticles were blended in the iron salts–CS gel solution. The amount of TiO2 varied from 5 mmol, 10 mmol to 15 mmol, which corresponds, respectively, to the molar ratios TiO2/Fe equal to 1, 2 and 3. By adding TiO2, the color of the solution changed from orange to white, as shown in Figure 7A. Then, the mixture was dropped through a syringe into the hardening sodium hydroxide solution to create spherical CS gel beads. The beads were washed several times with deionizer water to remove any residual alkali and dried in an oven at 50 °C.
As seen in Figure 7B, the obtained beads are black for those containing iron only, and they take on a greenish coloration in the presence of TiO2. All the prepared beads exhibited magnetic behavior in the presence of an external magnetic field (Figure 7C). The TiO2-CS beads were prepared following the same alkaline co-precipitation method without addition of iron salts.

3.3. CDM Removal Assays

The PEF process of CDM degradation was performed in a cyclic mode, as shown in Figure 8. The process was composed of two reactors connected in series, allowing the treatment of (400 mL, 30 mgL−1) CDM solution. A recirculation flow (200 mL min−1) was set using a pump to connect the two reactors.
The first reactor is electrochemical, composed of a glass beaker with a capacity of 250 mL, which permits the production of H2O2 by the reduction of dissolved oxygen on the cathode surface, which is a carbon felt (19.5 × 6 × 0.3 cm, Mersen, Barcelona, Spain) placed around the inner wall of the cylindrical cell, while the anode is a boron-doped diamond (BDD: 5 × 2.5 × 0.2 cm, Neocoat S.A.) placed in the middle of the cell. The system was run under a galvanostatic mode using an E 3512 A generator (Agilent, Santa Clara, CA, USA).
Na2SO4 was added previously as a support electrolyte at a concentration of 0.01 M, and the pH was adjusted to 3 using sulfuric acid solution. Air bubbling was maintained 15 min before the start of the reaction and through the treatment in order to saturate the medium with oxygen.
The second reactor is a cylindrical glass cell. A low-consumption UV-LED lamp (365 nm, 40 W, 550 lumens) from Luckylight electronics was placed above it, emitting at a wavelength equal to 365 nm. The catalyst was added into the photocatalytic reactor to promote its activation and avoid its electrochemical degradation, especially in the presence of a powerful oxidizing BDD anode in the first reactor.
To highlight the contribution of (photoelectrolysis + H2O2) and photocatalysis processes to the degradation of CDM, assays were conducted, respectively, without catalytic beads and in the absence of a current. Likewise, to evaluate the adsorption process contribution, another assay was conducted in the absence of UV irradiation and a current.

3.4. Analytical Methods

3.4.1. Determination of CDM Concentration

During experiments, the samples were collected and filtered prior to analysis through a 0.45 μm pore-size cellulose acetate membrane. CDM was quantified by HPLC (Agilent 1260 equipped with UV detector) with a C18 reverse-phase (4.6 mm × 250 mm, 5 μm; Agilent) column. The diode array detector was set at a fixed wavelength equal to 240 nm. The eluent was water/acetonitrile (60/40), with a flow rate of 1 mL min−1.

3.4.2. Determination of Carboxylic Acids Concentrations

To identify and quantify the carboxylic acids generated during electrolysis, an HPLC was used with a diode array detector fixed at 206 nm. A Rezex™ ROA-Organic Acid H+ (8%) (300 × 7.8 mm, i.e., 8 µm) column was used and placed in an oven at 60 °C. The eluent was 0.005 N H2SO4 solution pumped at a flow rate of 0.5 mL min−1. The identification of carboxylic acids was based on a comparison of the retention times with those of pure standards.

3.4.3. Determination of Ions Concentrations

The ions generated were quantitatively followed by an ion chromatography system DIONEX ICS-3000. The separation of ions was performed by a Metrosep A Supp 5 column (4.0 × 250 mm). The eluent was 3.2 mmol L−1 Na2CO3 and 1 mmol L−1 NaHCO3 at a flow rate of 0.7 mL min−1. The limit of quantification (LOQ) of all chromatographic methods was 0.1 mg L−1.

3.4.4. Total Organic Carbon Measurements

The total organic carbon (TOC) was measured via catalytic high-temperature combustion by multi N/C 3100 equipment (Analytic Jena, Germany) coupled with a non-dispersive infrared detector. The percentage removal of TOC was calculated using the following equation:
%   removal   of   TOC = TOC 0 TOC t TOC 0 × 100  
with TOC0 and TOCt representing the initial TOC and that at instant t.

3.4.5. Determination of Fe Concentration

The Fe concentration was determined by Inductively Coupled Plasma ICP (model: Optima 4300 DV Perkin Elmer Instruments). To obtain a metal solution from the heterogeneous catalyst, an acid digestion method was carried out using a concentrated nitric acid solution with the beads and placed on an autoclave at 121°C (Danish standard DS250).

3.4.6. Determination of H2O2 Concentration

The concentration of H2O2 was determined by a colorimetric method using a (Thermo Electron Corporation Helios) spectrophotometer because a titanium oxalate complexing agent can react with H2O2, producing a yellow peroxo–titanium complex, which absorbs at λmax = 400 nm [71].

3.4.7. Characterization of the Synthesized Catalysts

The surface morphology of the catalytic beads was observed using a scanning electron microscope (JEOL JSM-6700F). The crystalline structure of the obtained catalysts was determined by X-Ray Diffraction (XRD, X’Pert PRO MPD). Finally, the Physical Properties Measurement System equipment (PPMS Ever Cool-II 9T) with a Vibrating Sample Magnetometer (VSM) at 298 K was used to explore the magnetic properties of the catalysts.

3.5. Specific Energy Consumption

The electric energy consumption (EEC) per unit volume of treated solution (kWh m−3) was calculated according to the following equation [72]:
EEC ( kWh m 3 ) = I × E   × t V
where I is the current applied (A), E is the average cell voltage (V), t is the electrolysis time (h), and V is the solution volume (L or m3).

4. Conclusions

In summary, magnetic TiO2/Fe3O4-CS beads were synthesized following an easy in situ co-precipitation approach. The molar ratio TiO2/Fe was studied, and the enhanced catalytic activity of TiO2/Fe3O4-CS, with a molar ratio equal to 2, was probably due to the reduced recombination of charge carriers on the surface of the catalyst.
The performance of the PEF process using TiO2/Fe3O4-CS(2) as a photocatalyst for the treatment of real wastewater doped with CDM insecticide was evaluated under optimal experimental conditions. Complete CDM removal was attained in 1 h, and more than 80% TOC abatement was achieved after 6 h of treatment, with simple carboxylic acids as the main by-product.
The catalytic activity of TiO2/Fe3O4-CS(2) was satisfactorily validated in four consecutive cycles, and a slight decrease was obtained between the first and the fourth runs. Post-reaction catalyst characterization showed a high stability of magnetic properties despite the oxidation of Fe3O4 to Fe2O3, which is known for its good catalytic activity, sometimes similar to Fe3O4. Thus, the slight decrease in the catalytic activity could be attributed to the leaching of metals caused by the effect of the oxidizing conditions and UV-LED irradiation on catalytic beads.
This study offered a simple approach for constructing an eco-friendly, simple recovery and efficient bifunctional catalyst for advanced oxidation processes for treating recalcitrant organic pollutants in wastewater.

Author Contributions

Conceptualization, S.R.; Methodology, S.R. and A.M.D.; Validation, M.P. and M.Á.S.; Resources, M.P. and M.Á.S.; Writing—Original Draft Preparation, S.R.; Writing—Review and Editing, S.R., A.M.D., L.M., N.A., M.P. and M.Á.S.; Visualization, A.M.D., M.P. and M.Á.S.; Supervision, L.M., N.A., M.P. and M.Á.S.; Funding Acquisition, A.M.D., N.A., M.P. and M.Á.S. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Xunta de Galicia, grant number (ED481B 2019/091). This research was also funded by Project PID2020-113667GB-I00 464 funded by MCIN/AEI /10.13039/501100011033. The authors are grateful to Xunta de Galicia for funding the researcher Aida Maria Díez Sarabia (ED481B 2019/091). This work was also financially supported by the Research Unit on Electrochemistry, Materials and Environment (UR16ES02) at the University of Kairouan in Tunisia.

Data Availability Statement

Data is contained within the article.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Malakootian, M.; Shahesmaeili, M.A.; Fraji, M.; Amiri, H.; Silva Martinez, S. Advanced oxidation processes for the removal of organophosphorus pesticides in aqueous matrices: A systematic review and meta-analysis. Process. Saf. Environ. Prot. 2020, 134, 292–307. [Google Scholar] [CrossRef]
  2. Torres, N.H.; de Oliveira Santiago Santos, G.; Ferreira, L.F.R.; Américo-Pinheiro, J.H.P.; Eguiluz, K.I.B.; Salazar-Banda, G.R. Environmental aspects of hormones estriol, 17β-estradiol and 17α-ethinylestradiol: Electrochemical processes as next-generation technologies for their removal in water matrices. Chemosphere 2021, 267, 128888. [Google Scholar] [CrossRef] [PubMed]
  3. Trellu, C.; Olvera Vargas, H.; Mousset, E.; Oturan, N.; Oturan, M.A. Electrochemical technologies for the treatment of pesticides. Curr. Opin. Electrochem. 2021, 26, 100677. [Google Scholar] [CrossRef]
  4. Rajasekhar, B.; Venkateshwaran, U.; Durairaj, N.; Divyapriya, G.; Nambi, I.M.; Joseph, A. Comprehensive treatment of urban wastewaters using electrochemical advanced oxidation process. J. Environ. Manag. 2022, 266, 110469. [Google Scholar] [CrossRef] [PubMed]
  5. Heidari, Z.; Pelalak, R.; Alizadeh, R.; Oturan, N.; Shirazian, S.; et Oturan, M.A. Application of Mineral Iron-Based Natural Catalysts in Electro-Fenton Process: A Comparative Study. Catalysts 2021, 11, 1–18. [Google Scholar] [CrossRef]
  6. Ganzenko, O.; Trrellu, C.; Oturan, N.; Huguenot, D.; Péchaud, Y.; Van Hullebusch, E.D.; Oturan, M.A. Electro-Fenton treatment of a complex pharmaceutical mixture: Mineralization efficiency and biodegradability enhancement. Chemosphere 2020, 253, 126659. [Google Scholar] [CrossRef] [PubMed]
  7. Hussain, S.; Aneggi, E.; et 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]
  8. Guo, W.; Li, T.; Chen, Q.; Wan, J.; Zhang, J.; Wu, B.; Wang, Y. The roles of wavelength in the gaseous toluene removal with OH from UV activated Fenton reagent. Chemosphere 2021, 275, 129998. [Google Scholar] [CrossRef]
  9. Babuponnusami, A.; Muthukumar, K. Advanced oxidation of phenol: A comparison between Fenton, electro-Fenton, sono-electro-Fenton and photo-electro-Fenton processes. J. Chem. Eng. 2012, 183, 1–9. [Google Scholar] [CrossRef]
  10. Babuponnusami, A.; Muthukumar, K. Removal of phenol by heterogenous photo electro Fenton-like process using nano-zero valent iron. Sep. Purif. Technol. 2012, 98, 130–135. [Google Scholar] [CrossRef]
  11. Chu, A.C.; Sahu, R.S.; Chou, T.H.; Shih, Y. Magnetic Fe3O4@TiO2 nanocomposites to degrade bisphenol A, one emerging contaminant, under visible and long wavelength UV light irradiation. J. Environ. Chem. Eng. 2021, 9, 105539. [Google Scholar] [CrossRef]
  12. Sun, Q.; Hong, Y.; Liu, Q.; Dong, L. Synergistic operation of photocatalytic degradation and Fenton process by magnetic Fe3O4 loaded TiO2. Appl. Surf. Sci. 2018, 430, 399–406. [Google Scholar] [CrossRef]
  13. Mukhtar, F.; Munawar, T.; Nadeem, M.S.; Rehman, M.N.; Khan, S.A.; Koc, M.; Batool, S.; Hasan, M.; Iqbal, F. Dual Z-scheme core-shell PANI-CeO2-Fe2O3-NiO heterostructured nanocomposite for dyes remediation under sunlight and bacterial disinfection. Environ. Res. 2022, 215, 114140. [Google Scholar] [CrossRef] [PubMed]
  14. Mukhtar, F.; Munawar, T.; Nadeem, M.S.; Khan, S.A.; Koc, M.; Batool, S.; Hasan, M.; Iqbal, F. Enhanced sunlight-absorption of Fe2O3 covered by PANI for the photodegradation of organic pollutants and antimicrobial inactivation. Adv. Powder Technol. 2022, 33, 103708. [Google Scholar] [CrossRef]
  15. Goswami, A.; Jiang, J.-Q.; Petri, M. Treatability of five micro-pollutants using modified Fenton reaction catalyzed by zero-valent iron powder (Fe(0)). J. Environ. Chem. Eng. 2021, 9, 105393. [Google Scholar] [CrossRef]
  16. Wang, Y.; Fang, J.; Crittenden, J.C.; Shen, C. Novel RGO/α-FeOOH supported catalyst for Fenton oxidation of phenol at a wide pH range using solar-light-driven irradiation. J. Hazard. Mater. 2017, 239, 321–329. [Google Scholar] [CrossRef]
  17. Luo, H.; Zeng, Y.; He, D.; Pan, X. Application of iron-based materials in heterogeneous advanced oxidation processes for wastewater treatment: A review. J. Chem. Eng. 2021, 407, 127191. [Google Scholar] [CrossRef]
  18. Dudchenko, N.; Pawar, S.; Perelshtein, I.; Fixler, D. Magnetite Nanoparticles: Synthesis and Applications in Optics and Nanophotonics. Materials 2022, 15, 2601. [Google Scholar] [CrossRef]
  19. Nguyen, M.D.; Tran, H.-V.; Xu, S.; Lee, T.R. Fe3O4 Nanoparticles: Structures, Synthesis, Magnetic Properties, Surface Functionalization, and Emerging Applications. Appl. Sci. 2021, 11, 11301. [Google Scholar] [CrossRef]
  20. Gopinath, K.P.; Madhav, V.M.; Krishnan, A.; Malolan, R.; Rangarajan, G. Present applications of titanium dioxide for the photocatalytic removal of pollutants from water: A review. J. Environ. Manag. 2020, 270, 110906. [Google Scholar] [CrossRef]
  21. Mukhtar, F.; Munawar, T.; Nadeem, M.S.; Rehman, M.N.; Batool, S.; Hasan, M.; Riaz, M.; Rehman, K.; Iqbal, F. Highly efficient tri-phase TiO2-Y2O3-V2O5 nanaocomposite: Structural, optical, photocatalyst and antibacterial studies. J. Nanostruct. Chem. 2022, 12, 547–564. [Google Scholar] [CrossRef]
  22. Liu, C.; Dai, H.; Tan, C.; Pan, Q.; Hu, F.; Peng, X. Photo-Fenton degradation of tetracycline over Z-sheme Fe-g-C3N4/Bi2WO6 heterojunctions: Mechanism insight, degradation pathways and DFT calculation. Appl. Catal. B 2022, 310, 121326. [Google Scholar] [CrossRef]
  23. Sun, L.; Zhou, Q.; Mao, J.; Ouyang, X.; Yuan, Z.; Song, X.; Gong, W.; Mei, S.; Xu, W. Study on Photocatalytic Degradation of Acid Red 73 by Fe3O4@TiO2 Exposed (001) Facets. Appl. Sci. 2022, 12, 3574. [Google Scholar] [CrossRef]
  24. Rezgui, S.; Díez, A.M.; Monser, L.; Adhoum, N.; Pazos, M.; et Sanromán, M.A. ZnFe2O4-chitosan magnetic beads for the removal of chlordimeform by photo-Fenton process under UVC irradiation. J. Environ. Manag. 2021, 283, 111987. [Google Scholar] [CrossRef] [PubMed]
  25. Rezgui, S.; Amrane, A.; Fourcade, F.; Assadi, A.; Monser, L.; Adhoum, N. Electro-Fenton catalyzed with magnetic chitosan beads for the removal of Chlordimeform insecticide. Appl. Catal. B 2018, 226, 346–359. [Google Scholar] [CrossRef]
  26. Lee, M.; Chen, B.-Y.; Den., W. Chitosan as a Natural Polymer for Heterogeneous Catalysts Support: A. Short Review on Its Applications. Appl. Sci. 2015, 5, 1272–1283. [Google Scholar] [CrossRef] [Green Version]
  27. Zhang, W.; Jia, S.; Wu, Q.; Wu, S.; Ran, J.; Lui, Y.; Hou, J. Studies of the magnetic field intensity on the synthesis of chitosan-coated magnetite nanocomposites by co-precipitation method. Mater. Sci. Eng. C 2012, 32, 381–384. [Google Scholar] [CrossRef]
  28. Donadel, K.; Felisberto, M.D.V.; Fávere, V.T.; Rigoni, M.; Batistela, N.J.; Laranjeira, C.M.C. Synthesis and characterization of the iron oxide magnetic particles coated with chitosan biopolymer. Mater. Sci. Eng. C 2008, 28, 509–514. [Google Scholar] [CrossRef]
  29. Moura, F.C.C.; Araujo, M.H.; Costa, R.C.C.; Ardisson, J.D.; Macedo, W.A.A.; Lago, R.M. Efficient use of Fe metal as an electron transfer agent in a heterogeneous Fenton system based on Fe0/Fe3O4 composites. Chemosphere 2005, 60, 1118–1123. [Google Scholar] [CrossRef]
  30. Sun, Q.; Leng, W.; Li, Z.; Xu, Y. Effect of surface Fe2O3 clusters on the photocatalytic activity of TiO2 for phenol degradation in water. J. Hazard. Mater. 2012, 229–230, 224–232. [Google Scholar] [CrossRef]
  31. Afzal, S.; Julkapli, N.M.; Mun, L.K. Visible light active TiO2/CS/Fe3O4 for nitrophenol degradation: Studying impact of TiO2, CS and Fe3O4 loading on the optical and photocatalytic performance of nanocomposite. Mater. Sci. Semicond. Process. 2021, 131, 105891. [Google Scholar] [CrossRef]
  32. Díez, A.M.; Pazos, M.; Sanromán, M.A. Synthesis of magnetic-photo-Fenton catalyst for degradation of emerging pollutant. Catal. Today 2019, 328, 267–273. [Google Scholar] [CrossRef]
  33. Li, Y.; Cheng, H. Chemical kinetic modeling of organic pollutant degradation in Fenton and solar photo-Fenton processes. J. Taiwan Inst. Chem. Eng. 2021, 123, 175–184. [Google Scholar] [CrossRef]
  34. Burbano, A.A.; Dionysiou, D.D.; Suidan, M.T.; Richardson, T.L. Oxidation kinetics and effect of pH on the degradation of MTBE with Fenton reagent. Water Res. 2016, 39, 107–118. [Google Scholar] [CrossRef] [PubMed]
  35. Molamahmood, H.V.; Geng, W.; Wei, Y.; Miao, J.; Yu, S.; Shahi, A.; Chen, C.; Long, M. Catalyzed H2O2 decomposition over iron oxides and oxyhydroxides: Insights from oxygen production and organic degradation. Chemosphere 2022, 291, 133037. [Google Scholar] [CrossRef]
  36. Li, Q.; Kong, H.; Jia, R.; Shao, J.; He, Y. Enhanced catalytic degradation of amoxicillin with TiO2–Fe3O4 composites via a submerged magnetic separation membrane photocatalytic reactor (SMSMPR). RSC Adv. 2019, 9, 12538–12546. [Google Scholar] [CrossRef] [Green Version]
  37. Li, Q.; Kong, H.; Li, P.; Shao, J.; He, Y. Photo-Fenton degradation of amoxicillin via magnetic TiO2-graphene oxide-Fe3O4 composite with a submerged magnetic separation membrane photocatalytic reactor (SMSMPR). J. Hazard. Mater. 2019, 373, 437–446. [Google Scholar] [CrossRef]
  38. Bai, X.; Lyu, L.; Ma, W.; Ye, Z. Heterogeneous UV/Fenton degradation of bisphenol A catalyzed by synergistic effects of FeCo2O4/TiO2/GO. Environ. Sci. Pollut. Res. 2016, 23, 22734–22743. [Google Scholar] [CrossRef]
  39. Nwe, T.S.; Sikong, L.; Kokoo, R.; Khangkhamano, M. Photocatalytic activity enhancement of Dy-doped TiO2 nanoparticles hybrid with TiO2 (B) nanobelts under UV and fluorescence irradiation. Curr. Appl. Phys. 2016, 20, 249–254. [Google Scholar] [CrossRef]
  40. Nam, Y.; Lim, J.H.; Ko, K.C.; Lee, J.K. Photocatalytic activity of TiO2 nanoparticles: A theoretical aspect. J. Mater. Chem. A 2019, 7, 13833–13859. [Google Scholar] [CrossRef]
  41. Enesca, A.; Isac, L. The Influence of Light Irradiation on the Photocatalytic Degradation of Organic Pollutants. Materials 2020, 11, 2494. [Google Scholar] [CrossRef] [PubMed]
  42. Boruah, P.K.; Das, M.R. Dual responsive magnetic Fe3O4-TiO2/graphene nanocomposite as an artificial nanozymes for the colorimetric detection and photodegradation of pesticide in an aqueous medium. J. Hazard. Mater. 2020, 385, 121516. [Google Scholar] [CrossRef] [PubMed]
  43. Pourzad, A.; Sobhi, H.R.; Behbahani, M.; Esrafili, A.; Kalantary, R.R.; Kermani, M. Efficient visible light-induced photocatalytic removal of paraquat using N-doped TiO2@SiO2@Fe3O4 nanocomposite. J. Mol. Liq. 2020, 299, 112467. [Google Scholar] [CrossRef]
  44. Zahedifar, M.; Seyedi, N. Bare 3D-TiO2/magnetic biochar dots (3D-TiO2/BCDs MNPs): Highly efficient recyclable photocatalyst for diazinon degradation under sunlight irradiation. Phys. E Low-Dimens. Syst. Nanostruct. 2022, 139, 115151. [Google Scholar] [CrossRef]
  45. Hua, Y.; Wang, S.; Xiao, J.; Cui, C.; Wang, C. Preparation and characterization of Fe3O4/gallic acid/graphene oxide magnetic nanocomposites as highly efficient Fenton catalysts. RSC Adv. 2017, 7, 28979–28986. [Google Scholar] [CrossRef] [Green Version]
  46. Kasiri, M.B.; Aleboyeh, H.; Aleboyeh, A. Degradation of Acid Blue 74 using Fe-ZSM5 zeolite as a heterogeneous photo-Fenton catalyst. Appl. Catal. B 2008, 84, 9–15. [Google Scholar] [CrossRef]
  47. Tekbaş, M.; Yatmaz, H.C.; Bektaş, N. Heterogeneous photo-Fenton oxidation of reactive azo dye solutions using iron exchanged zeolite as a catalyst. Microporous Mesoporous Mater. 2008, 115, 594–602. [Google Scholar] [CrossRef]
  48. Jiang, B.; Niu, Q.; Li, C.; Oturan, N.; Oturan, M.A. Outstanding performance of electro-Fenton process for efficient decontamination of Cr(III) complexes via alkaline precipitation with no accumulation of Cr(VI): Important roles of iron species. Appl. Catal. B 2020, 272, 119002. [Google Scholar] [CrossRef]
  49. Hammouda, S.B.; Amrane, A.; Fourcade, F.; Assadi, A.; Adhoum, N.; Monser, L. Effective heterogeneous electro-Fenton process for the degradation of a malodorous compound indole using iron loaded alginate beads as a reusable catalyst. Appl. Catal. B 2016, 182, 47–58. [Google Scholar] [CrossRef]
  50. Liu, D.; Zhang, H.; Wei, Y.; Liu, B.; Lin, Y.; Li, G.; Zhang, F. Enhanced degradation of ibuprofen by heterogeneous electro-Fenton at circumneutral pH. Chemosphere 2018, 209, 998–1006. [Google Scholar] [CrossRef]
  51. Chmayssem, A.; Taha, S.; Hauchard, D. Scaled-up electrochemical reactor with a fixed bed three-dimensional cathode for electro-Fenton process: Application to the treatment of bisphenol A. Electrochim. Acta 2017, 225, 435–442. [Google Scholar] [CrossRef]
  52. Nidheesh, P.V.; Gandhimathi, R. Trends in electro-Fenton process for water and wastewater treatment: An overview. Desalination 2012, 299, 1–15. [Google Scholar] [CrossRef]
  53. Kang, B.; Dai, Y.; Zhang, H.; Chen, D. Synergetic degradation of chitosan with gamma radiation and hydrogen peroxide. Polym. Degrad. Stab. 2007, 92, 359–362. [Google Scholar] [CrossRef]
  54. Wang, S.-M.; Huang, Q.-Z.; Wang, Q.-S. Study on the synergetic degradation of chitosan with ultraviolet light and hydrogen peroxide. Carbohydr. Res. 2005, 340, 1143–1147. [Google Scholar] [CrossRef]
  55. Zazouli, M.A.; Ghanbari, F.; Yousefi, M.; Madihi-Bidgoli, S. Photocatalytic degradation of food dye by Fe3O4–TiO2 nanoparticles in presence of peroxymonosulfate: The effect of UV sources. J. Environ. Chem. Eng. 2017, 5, 2459–2468. [Google Scholar] [CrossRef]
  56. Maktabifard, M.; Zaborowska, E.; Makinia, J. Achieving energy neutrality in wastewater treatment plants through energy savings and enhancing renewable energy production. Rev. Environ. Sci. Biotechnol. 2018, 17, 655–689. [Google Scholar] [CrossRef] [Green Version]
  57. Ghazouani, M.; Bousselmi, L.; Akrout, H. Combined electrocoagulation and electrochemical treatment on BDD electrodes for simultaneous removal of nitrates and phosphates. J. Environ. Chem. Eng. 2020, 8, 104509. [Google Scholar] [CrossRef]
  58. Ghanbari, F.; Moradi, M.A. comparative study of electrocoagulation, electrochemical Fenton, electro-Fenton and peroxi-coagulation for decolorization of real textile wastewater: Electrical energy consumption and biodegradability improvement. J. Environ. Chem. Eng. 2015, 3, 499–506. [Google Scholar] [CrossRef]
  59. Gerek, E.E.; Yılmaz, S.; Koparal, A.S.; Gerek, Ö.N. Combined energy and removal efficiency of electrochemical wastewater treatment for leather industry. J. Water Process. Eng. 2019, 30, 100382. [Google Scholar] [CrossRef]
  60. Flores, N.; Sirés, I.; Garrido, J.A.; Centellas, F.; Rodriguez, R.M.; Cabot, P.L.; Brillas., E. Degradation of trans-ferulic acid in acidic aqueous medium by anodic oxidation, electro-Fenton and photoelectro-Fenton. J. Hazard. Mater. 2016, 319, 3–12. [Google Scholar] [CrossRef]
  61. Hammami, S.; Oturan, N.; Bellakhal, N.; Dachraoui, M.; Oturan, M.A. Oxidative degradation of direct orange 61 by electro-Fenton process using a carbon felt electrode: Application of the experimental design methodology. J. Electroanal. Chem. 2007, 610, 75–84. [Google Scholar] [CrossRef]
  62. Aplin, R.; Feitz, A.J.; Waite, T.D. Effect of Fe(III)-ligand properties on effectiveness of modified photo-Fenton processes. Water Sci. Technol. 2001, 44, 23–30. [Google Scholar] [CrossRef] [PubMed]
  63. Zazou, H.; Oturan, N.; Zhang, H.; Hamdani, M.; Oturan, M.A. Comparative study of electrochemical oxidation of herbicide 2,4,5-T: Kinetics, parametric optimization and mineralization pathway. Sustain. Environ. Res. 2017, 27, 15–23. [Google Scholar] [CrossRef]
  64. Mhemdi, A.; Oturan, M.A.; Oturan, N.; Abdelhédi, R.; Ammar, S. Electrochemical advanced oxidation of 2-chlorobenzoic acid using BDD or Pt anode and carbon felt cathode. J. Electroanal. Chem. 2013, 709, 111–117. [Google Scholar] [CrossRef]
  65. Skoumal, M.; Arias, C.; Cabot, P.L.; Centellas, F.; Garrido, J.A.; Rodriguez, R.M.; Brillas, E. Mineralization of the biocide chloroxylenol by electrochemical advanced oxidation processes. Chemosphere 2008, 71, 1718–1729. [Google Scholar] [CrossRef]
  66. Arai, M.; Zhao, F. Metal Catalysts Recycling and Heterogeneous/Homogeneous Catalysis. Catalysts 2015, 5, 868–870. [Google Scholar] [CrossRef] [Green Version]
  67. Thomas, N.; Dionysiou, D.D.; Pillai, S.C. Heterogeneous Fenton catalysts: A review of recent advances. J. Hazard. Mater. 2021, 404, 124082. [Google Scholar] [CrossRef]
  68. Rusevova, K.; Kopinke, F.-D.; Georgi, A. Nano-sized magnetic iron oxides as catalysts for heterogeneous Fenton-like reactions—Influence of Fe(II)/Fe(III) ratio on catalytic performance. J. Hazard. Mater. 2012, 241–242, 433–440. [Google Scholar] [CrossRef]
  69. Gijs, M.A.M. Magnetic bead handling on-chip: New opportunities for analytical applications. Microfluid. Nanofluid. 2004, 1, 22–40. [Google Scholar] [CrossRef] [Green Version]
  70. Li, Y.; Zhang, M.; Guo, M.; Wang, X. Preparation and properties of a nano TiO2/Fe3O4 composite superparamagnetic photocatalyst. Rare Metals 2009, 28, 423–427. [Google Scholar] [CrossRef]
  71. Almuaibed, A.M.; Townshend, A. Flow spectrophotometric method for determination of hydrogen peroxide using a cation exchanger for preconcentration. Anal. Chim. Acta 1994, 295, 159–163. [Google Scholar] [CrossRef]
  72. Martínez-Huitle, C.A.; Brillas, E. Decontamination of wastewaters containing synthetic organic dyes by electrochemical methods: A general review. Appl. Catal. B 2009, 87, 105–145. [Google Scholar] [CrossRef]
Figure 1. Performance comparison of several processes and effect of the addition of TiO2 to magnetic chitosan beads: (a,d) % removal of CDM; (b) −Ln (C/C0) vs. time; (c) [H2O2] vs. time: I = 50 mA in the presence of 1 g   L 1 of catalyst and at pHi = 3.
Figure 1. Performance comparison of several processes and effect of the addition of TiO2 to magnetic chitosan beads: (a,d) % removal of CDM; (b) −Ln (C/C0) vs. time; (c) [H2O2] vs. time: I = 50 mA in the presence of 1 g   L 1 of catalyst and at pHi = 3.
Catalysts 12 01425 g001
Figure 2. (a) Effect of catalyst dosage, (b) Effect of current intensity.
Figure 2. (a) Effect of catalyst dosage, (b) Effect of current intensity.
Catalysts 12 01425 g002
Figure 3. Treatment of a real effluent: (a) % removal of CDM; (b) % removal of TOC; (c) [carboxylic acids] vs. time (d) [ions] vs. time: I = 70 mA in the presence of 1 g   L 1 of TiO2/Fe3O4-CS(2) and at pHinitial = 3.
Figure 3. Treatment of a real effluent: (a) % removal of CDM; (b) % removal of TOC; (c) [carboxylic acids] vs. time (d) [ions] vs. time: I = 70 mA in the presence of 1 g   L 1 of TiO2/Fe3O4-CS(2) and at pHinitial = 3.
Catalysts 12 01425 g003
Figure 4. Stability of the catalytic activity of TiO2/Fe3O4-CS(2) beads.
Figure 4. Stability of the catalytic activity of TiO2/Fe3O4-CS(2) beads.
Catalysts 12 01425 g004
Figure 5. Morphology of catalyst: (a) TiO2/Fe3O4-CS(2) before use, (b) TiO2/Fe3O4-CS(2) after use.
Figure 5. Morphology of catalyst: (a) TiO2/Fe3O4-CS(2) before use, (b) TiO2/Fe3O4-CS(2) after use.
Catalysts 12 01425 g005
Figure 6. (a) X-ray diffraction patterns of TiO2/Fe3O4-CS(2) beads, (b) Magnetization curves of the fresh and used TiO2/Fe3O4-CS(2) beads.
Figure 6. (a) X-ray diffraction patterns of TiO2/Fe3O4-CS(2) beads, (b) Magnetization curves of the fresh and used TiO2/Fe3O4-CS(2) beads.
Catalysts 12 01425 g006
Figure 7. (A) Chitosan gel/Fe3O solutions. (B) Wet chitosan beads at a molar ratio TiO2/Fe: (1) 0; (2) 1; (3) 2 and (4) 3. (C) Superparamagnetic behavior of the TiO24-CS beads at a molar ratio TiO2/Fe = 2.
Figure 7. (A) Chitosan gel/Fe3O solutions. (B) Wet chitosan beads at a molar ratio TiO2/Fe: (1) 0; (2) 1; (3) 2 and (4) 3. (C) Superparamagnetic behavior of the TiO24-CS beads at a molar ratio TiO2/Fe = 2.
Catalysts 12 01425 g007
Figure 8. Cyclic electro-photocatalytic reactors: (I) electrochemical reactor: (1) power supply, (2) carbon felt electrode, (3) BDD electrode, (4) air supply, (5) magnetic stirrer, (6) peristaltic pump; (II) photocatalytic reactor: (5) magnetic stirrer, (7) UV-LED lamp, (8) catalytic beads.
Figure 8. Cyclic electro-photocatalytic reactors: (I) electrochemical reactor: (1) power supply, (2) carbon felt electrode, (3) BDD electrode, (4) air supply, (5) magnetic stirrer, (6) peristaltic pump; (II) photocatalytic reactor: (5) magnetic stirrer, (7) UV-LED lamp, (8) catalytic beads.
Catalysts 12 01425 g008
Table 1. Summary table of mineralization rates.
Table 1. Summary table of mineralization rates.
Parameter AssayRate of TOC Removal after 4 h of Treatment
Performance comparison of several processes TiO2/Fe3O4-CS(2) + UV-LED4.4 ± 1.6
H2O2 + UV-LED (Without catalyst)18.2 ± 3.3
TiO2/Fe3O4-CS(2) + H2O2 + UV-LED69.6 ± 2.7
Influence of TiO2 in magnetic chitosan beads Fe3O4-CS + H2O2 + UV-LED 33.9 ± 4.1
TiO2/Fe3O4-CS(2) + H2O2 + UV-LED69.6 ± 2.7
Effect of TiO2 loading content into magnetic chitosan beadsMolar ratio TiO2/Fe = 138.2 ± 3.5
Molar ratio TiO2/Fe = 269.6 ± 2.7
Molar ratio TiO2/Fe = 354.2 ± 5.6
Catalytic beads of TiO224.5 ± 1.9
Effect of catalyst dosage0.25 g L−140.9 ± 4.1
0. 5 g L−160.8 ± 3.2
1 gL−169.6 ± 2.7
1.5 g L−170.2 ± 1.7
2 g L−158.2 ± 6.2
Effect of current intensity50 mA68.6 ± 4.9
70 mA 76.9 ± 3.7
100 mA77.7 ± 1.2
Table 2. The comparison of photodegradation efficiency of TiO2/Fe3O4-CS catalytic beads with other TiO2/Fe3O4-based materials against different pesticides.
Table 2. The comparison of photodegradation efficiency of TiO2/Fe3O4-CS catalytic beads with other TiO2/Fe3O4-based materials against different pesticides.
CatalystsPesticidesDegradation Efficiency (%)Refs.
Fe3O4-TiO2/reduced graphene oxideAtrazine 99% within 40 min[42]
N-doped TiO2@SiO2@Fe3O4 nanocompositeParaquat 98.7% within 180 min[43]
Bare 3D-TiO2/magnetic biochar dotsDiazinon 98.5% within 30 min[44]
TiO2/Fe3O4-CSChlordimeform100% removal within 60 minPresent work
Table 3. Physicochemical properties of the collected wastewater.
Table 3. Physicochemical properties of the collected wastewater.
ParameterValue
Total organic carbon TOC (mg L−1)52.7
Chemical oxygen demand DCO (mg L−1)35
Biological oxygen demand BOD5 (mg L−1)2
Conductivity (mS cm−1)3.33
pH7.31
Turbidity (mg L−1)12
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Rezgui, S.; Díez, A.M.; Monser, L.; Adhoum, N.; Pazos, M.; Sanromán, M.Á. Magnetic TiO2/Fe3O4-Chitosan Beads: A Highly Efficient and Reusable Catalyst for Photo-Electro-Fenton Process. Catalysts 2022, 12, 1425. https://doi.org/10.3390/catal12111425

AMA Style

Rezgui S, Díez AM, Monser L, Adhoum N, Pazos M, Sanromán MÁ. Magnetic TiO2/Fe3O4-Chitosan Beads: A Highly Efficient and Reusable Catalyst for Photo-Electro-Fenton Process. Catalysts. 2022; 12(11):1425. https://doi.org/10.3390/catal12111425

Chicago/Turabian Style

Rezgui, Soumaya, Aida M. Díez, Lotfi Monser, Nafaa Adhoum, Marta Pazos, and M. Ángeles Sanromán. 2022. "Magnetic TiO2/Fe3O4-Chitosan Beads: A Highly Efficient and Reusable Catalyst for Photo-Electro-Fenton Process" Catalysts 12, no. 11: 1425. https://doi.org/10.3390/catal12111425

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