Next Article in Journal
A Review on Per- and Polyfluoroalkyl Substances in Pregnant Women: Maternal Exposure, Placental Transfer, and Relevant Model Simulation
Next Article in Special Issue
A Review on the Use of Metal Oxide-Based Nanocomposites for the Remediation of Organics-Contaminated Water via Photocatalysis: Fundamentals, Bibliometric Study and Recent Advances
Previous Article in Journal
Evaluating the Impact of Individual and Combined Toxicity of Imidacloprid, Cycloxaprid, and Tebuconazole on Daphnia magna
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Synergetic Photocatalytic Peroxymonosulfate Oxidation of Benzotriazole by Copper Ferrite Spinel: Factors and Mechanism Analysis

1
Department of Environmental Health Engineering, Faculty of Health, Zabol University of Medical Sciences, Zabol 9861615881, Iran
2
School of Environmental Studies, China University of Geosciences, Wuhan 430074, China
3
Unidad Docente Ingeniería Sanitaria, Departamento de Ingeniería Civil: Hidráulica, Energía y Medio Ambiente, E.T.S. de Ingenieros de Caminos, Canales y Puertos, Universidad Politécnica de Madrid, s/n, 28040 Madrid, Spain
4
Institute for Nanotechnology and Water Sustainability, College of Science, Engineering and Technology, University of South Africa, Johannesburg 1709, Florida, South Africa
5
Research Center for Health, Safety and Environment, Alborz University of Medical Sciences, Karaj 3149779453, Iran
6
Department of Environmental Health Engineering, Alborz University of Medical Sciences, Karaj 3149779453, Iran
*
Author to whom correspondence should be addressed.
Toxics 2023, 11(5), 429; https://doi.org/10.3390/toxics11050429
Submission received: 29 March 2023 / Revised: 24 April 2023 / Accepted: 28 April 2023 / Published: 4 May 2023
(This article belongs to the Special Issue Advanced Oxidation Processes and Biodegradability)

Abstract

:
The development of oxidation processes with the efficient generation of powerful radicals is the most interesting and thought-provoking dimension of peroxymonosulfate (PMS) activation. This study reports the successful preparation of a magnetic spinel of CuFe2O4 using a facile, non-toxic, and cost-efficient co-precipitation method. The prepared material exhibited a synergetic effect with photocatalytic PMS oxidation, which was effective in degrading the recalcitrant benzotriazole (BTA). Moreover, central composite design (CCD) analysis confirmed that the highest BTA degradation rate reached 81.4% after 70 min of irradiation time under the optimum operating conditions of CuFe2O4 = 0.4 g L−1, PMS = 2 mM, and BTA = 20 mg L−1. Furthermore, the active species capture experiments conducted in this study revealed the influence of various species, including OH, SO4•−, O2•−, and h+ in the CuFe2O4/UV/PMS system. The results showed that SO4•− played a predominant role in BTA photodegradation. The combination of photocatalysis and PMS activation enhanced the consumption of metal ions in the redox cycle reactions, thus minimizing metal ion leaching. Additionally, this maintained the reusability of the catalyst with reasonable mineralization efficiency, which reached more than 40% total organic carbon removal after four batch experiments. The presence of common inorganic anions was found to have a retardant effect on BTA oxidation, with the order of retardation following: HCO3 > Cl > NO3 > SO42−. Overall, this work demonstrated a simple and environmentally benign strategy to exploit the synergy between the photocatalytic activity of CuFe2O4 and PMS activation for the treatment of wastewater contaminated with widely used industrial chemicals such as BTA.

1. Introduction

Benzotriazole (BTA) is an emerging contaminant originating from a wide range of industrial and domestic applications, such as the protection of alloys against corrosion, automotive cooling and aircraft de-icing fluids, enhancement of fastness to light in fabrics, antifreezes, and the production of various detergents. BTA is a cyclic compound that has a benzene ring that is fused into a C-C bond with three nitrogen atoms [1,2]. The presence of BTA has been reported in effluent-receiving water bodies (~7–18 μg/L) [3], seawater, and urban runoff as the major sources in the natural environment [4]. Generally, BTA has a low sorption affinity on organic matter, is extremely soluble in water, and is polar, which is responsible for its high environmental mobility [5]. Consequently, its persistence and toxicity in the aquatic environment make it less susceptible to chemical-biological treatment processes [4,5,6]. Moreover, TBA has been reported as a potential human carcinogen and is toxic to aquatic organisms. Therefore, there is an urgent need to explore efficient water treatment technologies for BTA mineralization.
Advanced oxidation processes (AOPs) have enormous potential for eliminating emerging and refractory pollutants using various reactive oxygen species (ROS) such as singlet oxygen (1O2) molecules, sulfate (SO4•−), hydroxyl (OH), and superoxide (O2•−) radicals. Several AOPs, including Fenton- and photo-Fenton-like systems [7], photocatalytic reactions [2], ozonation, and ultrasound-assisted oxidation [8] have been used for oxidative BTA elimination. Sulfate based-AOPs such as peroxymonosulfate (HSO5, PMS) and peroxydisulfate (S2O82−, PDS)-mediated oxidation have emerged as promising and advantageous AOPs, owing to their flexible operational conditions. For example, sulfate-based AOPs are applicable over a wide pH range, unlike the Fenton process, which is effective over a narrow pH range in the acidic region [9]. PMS is considered to be environmentally benign because of its less harmful by-products (SO42−) [10]. Furthermore, SO4•− has higher oxidizing efficiency (E0 = 2.5–3.1 V) vs. OH (E0 = 1.9–2.7 V) at neutral pH, better selectivity, and higher persistency in aqueous solutions (30–40 µs), allowing longer contact times with recalcitrant pollutants. Additionally, SO4•− is less prone to interfering with water constituents, which is ideal for water and wastewater treatment [11]. Therefore, SO4•− can overcome the unescapable deficiencies of conventional AOPs [12]. In particular, the activation of the PMS or PDS precursors by UV or ultrasound irradiation, base and transition metal catalysts ( such as Fe2+, Mn2+, Co2+, Cu2+, etc.), heat, and carbon-based materials can produce SO4•− via electron transfer, as shown in Equation (1) [13,14]. In this regard, homogeneous transition-metal-catalyzed PMS activation is often preferred due to its ease of initiation at room temperature and pressure, requiring low energy as shown in Equations (2) and (3) to form SO4•− as the predominant oxidizing species [15]:
S2O82−/HSO5 → SO4•− + (HSO5, SO42−, OH)
HSO5 + e → SO4•− + OH
HSO5 + e → SO42− + OH
However, the homogeneous transition metal ion-mediated PMS activation system results in secondary pollution due to the difficulty of recovering the potentially toxic transition metal ions [16]. Contrary to this, heterogeneous catalytic PMS activation offers several advantages, such as recoverability and structural stability, low toxicity of catalysts, and minimal secondary pollution [17]. Given the difficulty associated with recovering nanocatalysts, magnetic Fe-based catalysts have been investigated due to their easy separation under an external magnetic field [18,19]. The heterogeneous catalytic activation of PMS largely depends on the interaction between PMS and the catalyst. As a result, a large specific surface area, excellent catalytic activity, and a highly porous structure are needed so that the activation rate of PMS can be improved [20]. Copper ferrites (CuFe2O4), which possess intrinsic surface hydroxyl sites and ferromagnetic properties, have been extensively studied for their effective role in activating PMS for the removal of refractory organic pollutants in water [21,22]. Despite its relatively low catalytic effectiveness compared to CoFe2O4, the low toxicity of CuFe2O4 makes it an important PMS activator. Furthermore, a significant linear correlation between the degradation rate and quantity of surface hydroxyl sites has been demonstrated [23]. More specifically, CuF2O4 as a Cu–Fe mixed-metal catalyst endows the surface with synergistic redox reactions for Cu3+/Cu2+, Cu+/Cu2+, and Fe2+/Fe3+ in a tetrahedral and/or octahedral structure to effectively generate SO4•− radicals from PMS. The photocatalytic PMS activation by CuFe2O4 stems from its suitable light absorption, which provides the photosynergistic effect that is favorable for organic pollutants' degradation [22,24]. In this context, the photogenerated electrons are donated to the PMS electron acceptor (electron transfer), which cleaves the O-O bond in the PMS molecule into SO4•− and OH radicals [25]. Furthermore, the recombination phenomenon of photoinduced charges is suppressed through the consumption of photoinduced electrons by PMS, leading to an amplification of the catalytic efficiency [13,26]. In addition, ion leaching in aqueous environments is particularly minimized due to the strong interaction of bimetals and the convenient recovery of the catalyst after its application.
Therefore, this study leveraged the intrinsic properties of CuFe2O4, including its ability to absorb UV photons and catalyze PMS activation, to construct a CuFe2O4/UV/PMS reaction system. Such an integrated process of CuFe2O4/UV/PMS without sludge generation and secondary pollution is attractive for the fast and efficient removal of BTA from water. Several studies applied the CuFe2O4/PMS process and reported promising results on the catalytic oxidation of different micropollutants at neutral pH [27,28,29,30]. However, the exploitation of the well-integrated CuFe2O4/UV/PMS for the removal of BTA has been rarely reported. Therefore, we used a simple and environmentally friendly co-precipitation method to obtain CuFe2O4 nanoparticles and used it to activate PMS for BTA degradation under UV irradiation. The structural, morphological, and physicochemical properties of CuFe2O4 were investigated. We first studied the activity of the different reaction systems toward BTA degradation and found that the UV-assisted CuFe2O4/PMS combination transcended other combinations. Then, the interactive effect of operational parameters, including PMS dosage, catalyst dosage, different BTA concentrations, and reaction time, on the catalytic performance was optimized using response surface methodology (RSM) via central composite design (CCD). The reusability and stability of CuFe2O4 were consecutively analyzed through four cycles, and the treatment of BTA in the presence of interfering water constituents was evaluated under the optimum conditions. Finally, the probable transformation pathways of BTA were explored. This study provides a simple strategy for coupling UV and catalytic PMS activation over a magnetically separable catalyst to ensure high catalytic performance for the elimination of organic pollutants and easy recovery of the catalyst.

2. Experimental Section

2.1. Chemicals

Benzotriazole (C6H5N3, 99%), copper nitrate trihydrate (Cu(NO3)2·3H2O, 99%), and ferric nitrate nonahydrate (Fe(NO3)3·9H2O, 99%) were purchased from Merck Inc. Potassium peroxymonosulfate (Oxone, KHSO5·0.5KHSO4·0.5K2SO4), tert-butyl alcohol (TBA, (CH3)3COH), p-Benzoquinone (BQ, C6H4O2), methanol (MeOH, CH3OH), and potassium iodide (KI) were obtained from Sigma-Aldrich Inc., (St. Louis, MO, USA). All chemicals were used as received without further purification, and deionized water (DI-water) was used for all aqueous solutions.

2.2. Preparation of Catalyst

Magnetic CuFe2O4 nanoparticles were prepared using a co-precipitation method from the precursors Cu(NO3)2·3H2O and Fe(NO3)3·9H2O in a 1:2 molar ratio as previously reported [31].

2.3. Material Characterization

Powder X-ray diffraction (XRD) was performed on a Quantachrome/NOVA 2000X-ray diffractometer with Cu radiation (λ = 1.54 nm), operated at 30 mA and 40 kV in a 2θ range of 10° to 80°. Brunauer-Emmett-Teller analysis (BET, ASAP 2010; Micromeritics, Norcross, GA, USA) was applied to determine the surface area, average pore size, and volume. The magnetic features of the catalyst were studied using a vibrating sample magnetometer (VSM, Lakeshore 7400, Westerville, OH, USA). Meanwhile, the detailed microstructure of the catalyst was investigated using transmission electron microscopy (TEM, JEM-2100; Jeol; Akishima, Tokyo, Japan) with high resolution at 100 kV. A field emission scanning electron microscope (FESEM, Mira 3-XMU, Tescan USA Inc., Warrendale, PA, USA) coupled with an energy dispersive X-ray spectrometer (EDS) device was exploited for the morphology and qualitative chemical analysis of the catalyst. Optical response measurements were conducted using a UV–Vis diffuse reflectance spectrometer (DRS, Hitachi Ltd., Chiyoda, Tokyo, Japan) in the wavelength range of 190 to 800 nm.

2.4. Degradation Experiments and Sample Analyses

A cylindrical 300 mL quartz photoreactor equipped with a 6.0 W Hg UV-C lamp (Philips, The Netherlands) with a luminous intensity of 7800 cd and fitted with a 254 nm filter was used for the photocatalytic/PMS experiments at 25 ± 5 °C. The initial pH in all experiments (unless stated otherwise) was that of the BTA solution without adjustment. For each experiment, a certain mass of catalyst was put into the BPA solution (200 mL) and mechanically stirred (RW 20, IKA, Staufen im Breisgau, Germany) at 250 r/min for 60 min to ensure the adsorption-desorption equilibrium of the BTA molecules on the catalyst surface. Thereafter, a specific amount of PMS was added, and the light source was switched on to initiate the photocatalytic/PMS BTA oxidation reaction. At regular intervals, 2 mL of BTA solution was withdrawn, the catalyst was magnetically separated, and 1 mL of Na2S2O3. 5H2O (20 mM) was added before analysis using high-performance liquid chromatography (HPLC, KNUER, Berlin, Germany). The HPLC was equipped with a C18 separation column (100–5; 4.6 mm × 250 mm, 5 µm), a 2500 UV–visible detector, and BTA measurements were conducted at a maximum wavelength of 254 nm. The mobile phase was a 50:50% (v/v) water-acetonitrile mixture with a flow rate of 1.0 mL min−1. The degradation efficiency (%) was used to describe the performance of the process according to Equation (4):
D e g r a d a t i o n   e f f i c i e n c y   % = C 0 C t C 0 × 100 ( % )
where C0 and Ct are the BTA concentrations at the initial and certain treatment times, respectively. The photochemical stability of the catalyst was assessed in four repeated experiments. An atomic absorption spectrophotometer (AAS) (Analytikjena vario 6, Jena, Thueringen, Germany) was used to determine the concentrations of Fe and Cu in the filtrate for each cycle. Additionally, for each cycle, total organic carbon (TOC) determination was conducted to monitor the mineralization efficiency using a Shimadzu VCHS/CSN, Japan. Potassium iodide (KI), benzoquinone (BQ), tert-butyl alcohol (TBA), and methanol (MeOH) were applied as quenchers of the ROS. The effect of various anions existing in natural water matrices was studied during BTA removal.

2.5. Response Surface Method Experimental Design and Data Analysis

To determine the effect of the main factors on optimal degradation conditions of BTA, central composite design (CCD) was used under response surface methodology (RSM). In this experiment, the catalyst loading (x1), PMS dosage (x2), BTA concentration (x3), and irradiation time (x4) were varied at five various coded levels, as indicated in Table 1. The coded value of the ith variable (xi) was described by Equation (5):
x i = X i X 0 X i
where Xi is the actual value of the variables; X0 is Xi at the center point; and ∆Xi is the step with the maximum and minimum values of Xi. With the application of the CCD matrix, 30 tests (24 = 16 factorial points, 2 × 4 = 8 axial points, and 6 replications at the center points) were determined for four input variables (Table S1). The quadratic polynomial Equation 6 can be used to estimate the effects of four independent variables on the efficiency of the photocatalytic PMS activation process. Here, y is a predicted response variable of BTA degradation efficiency, b0 is the intercept, bi, bij, and bii are the linear, interaction, and quadratic regression coefficients, respectively, and ε is residual error.
y B T A % p r e d i c t e d = b 0 + b i x i + b i j x i x j + b i i x i 2 + ε
The significance and adequacy of the quadratic regression model were estimated by analysis of variance (ANOVA). The R-squared (R2) and the adjusted R2 (R2adj) values were the criteria for verifying the statistical significance of the second-order model.

3. Results and Discussion

3.1. CuFe2O4 Spinel Morphology and Microstructure Analysis

To investigate the microstructure and morphology of pure CuFe2O4, FESEM, and TEM techniques were conducted. Figure 1A shows a FESEM image of the catalyst, which confirms an irregular polygon morphology and a slightly agglomerated state emanating from the nanoscale size of the CuFe2O4 (<26 nm). Obviously, an uneven and rough surface was witnessed from the TEM image of the CuFe2O4 catalyst (Figure 1C,D), which is favorable for the diffusion of reactants (organic pollutant and oxidant) and increase interaction with active sites of the catalyst. Ultimately, this accelerates the redox reaction on the catalyst surface, leading to pollutant degradation. In addition, EDS analysis (Figure 1B) shows that the catalyst contained the elements Cu (25.98%), Fe (47.41%), and O (26.62%) in a molar ratio of nearly 1:1.82, which was almost equal to the molar ratio of the iron and copper precursors used in the preparation of the CuFe2O4 catalyst.

3.2. Textural and Surface Area, Magnetic and Optical Properties

The crystalline phase composition of CuFe2O4 was determined by XRD (Figure 2A). The main peaks observed at 2θ of 18.22°, 30.30°, 35.85°, 43.32°, 57.39°, and 62.87° were assigned to crystal planes of (1 1 1), (2 2 0), (3 1 1), (4 0 0), (5 1 1), and (4 4 0), which are characteristic of the spinel ferrite structure of CuFe2O4 (JCPDS no. 25-0283) [21]. As demonstrated by XRD patterns, the narrow and strong diffraction peaks are indicative of the high crystallinity of the catalyst, which could be expected to exhibit high activity [27]. Furthermore, the absence of additional XRD peaks is an indication of the phase purity of the prepared material. Accordingly, the Debye–Scherrer equation was used to calculate the crystalline size of CuFe2O4 nanoparticles. The equation is given as D = Kλ/βcosθ, where K is the Scherrer constant (value 0.94), λ is the X-ray wavelength (0.154 nm), β is the FWHM (full width at half maximum) of photocatalysts, D is the calculated crystalline size, and θ is the diffraction angle. Using this equation, the average crystallite size of CuFe2O4 nanoparticles was calculated at 16.7 nm.
The N2 adsorption–desorption isotherm plot (Figure 2B) and the pore size distribution (Figure 2C) were measured at −196.88 °C. The Langmuir isotherm is related to a type IV curve in the IUPAC classification with a hysteresis loop at the relative pressure (P/P0) range between 0.4 and 1. This confirms that the catalyst has a mesoporous structure. Accordingly, the BET surface area of the pure CuFe2O4 nanoparticles (insert in Figure 2B) was calculated to be 201.898 m2 g−1, which indicates a considerable surface area and significant adsorption capacity for the as-prepared catalyst. The total pore volume and micropore volume were 0.191 and 0.0126 mL g−1, respectively, and the average pore diameter was 3.7 nm. The pore size distribution in the range of 1–30 nm manifests a mesoporous structure with an average size of about 2.6 nm (Figure 2C). Therefore, the large BET surface area with many micropores and mesopores could provide a rich source of surface reaction sites to activate PMS and also adsorb BTA [32].
The VSM technique was applied in the presence of a magnetic field ranging from −40 to +40 kOe to study the magnetic properties of CuFe2O4 (Figure 2D). The saturation magnetization value (Ms) of CuFe2O4 was found to be 35.2 emu g−1. Therefore, it is expected that CuFe2O4 as an ideal ferromagnetism can be easily separated from the aqueous matrix within a short time using an external magnetic field. This ensures high-quality recycling and reuse of the nanocatalyst with minimum risks of secondary contamination of the environment and simplifies the practical application of these nanoparticles. The light absorption capacity of CuFe2O4 was investigated by ultraviolet–visible (UV–Vis) diffuse reflection spectroscopy (DRS). Figure 2E depicts that CuFe2O4 has a strong photoresponse in the UV region extending to the visible region, making it an ideal catalyst for UV–Vis-mediated PMS activation. Furthermore, this wide photoresponse in the visible region is also important in practical applications where solar light could be utilized as the activation energy for the catalyst.

3.3. Performance Evaluation towards BTA Oxidation

Figure 3 describes the BTA removal profiles of various catalytic systems under controlled conditions. Notably, in the presence of PMS and UV light separately, only 9.9 and 7.6% of BTA were degraded within 60 min of reaction, respectively, showing a low contribution of direct oxidation in BTA degradation under ambient conditions. Moreover, this is an indication of the stability of BTA towards photolysis, which could be responsible for its persistence in the environment under solar exposure. Additionally, the poor reactivity of the persulfate ions towards BTA oxidation was evident, hence the need for activation. In contrast, the photolysis of PMS improved the oxidation system, with BTA degradation reaching approximately 29.4% after 60 min of UV irradiation. Specifically, UV irradiation triggered the activation of PMS to produce SO4•− and OH radicals through the cleavage of the peroxide bond of PMS. This results in boosted BTA degradation compared to UV photolysis and direct persulfate ion oxidation. Meanwhile, the presence of CuFe2O4 spinel (0.2 g L−1) promoted BTA removal up to 31% in 60 min. This indicates that the redox capacity and adsorption effect of CuFe2O4 nanoparticles towards BTA molecules provided the necessary catalytic sites for contact with BTA molecules, leading to their degradation [26]. Therefore, the adsorption-desorption equilibrium was established and fixed at 60 min before irradiation for all the tested samples. As expected, BTA degradation was further boosted over CuFe2O4/UV and CuFe2O4/PMS coupled systems, reaching 53.6% and 62.2% BTA removal, respectively, demonstrating the catalytic ability of CuFe2O4 for the activation of PMS and the UV light promoting photocatalytic activity of the material. The synergistic redox action of Cu and Fe sites on the CuFe2O4 surface induced the PMS decomposition to produce large amounts of SO4•− radicals in the CuFe2O4/PMS oxidation system compared to the individual systems. Similarly, coupling CuFe2O4 with UV irradiation initiated activation of the photocatalyst to form various reactive species, which collectively contributed to improved photocatalytic degradation of BTA. Despite the improved BTA degradation observed in the various binary systems, the greatest synergistic effect was observed in the binary oxidation system coupled with UV light irradiation. Specifically, the CuFe2O4/UV/PMS system attained up to 73.1% BTA degradation in 60 min, thus confirming the effectiveness of the photocatalytic PMS oxidation. The enhanced performance of the CuFe2O4/UV/PMS system can be linked to the collective contribution of photocatalysis and photocatalytic activation of PMS, which yielded powerful SO4•− and •OH radicals, responsible for BTA mineralization [33]. Furthermore, it can be inferred that the higher surface area observed in BET analysis coincided with the higher mass transfer of BTA molecules, leading to accelerated reaction rates and mineralization of BTA molecules [9,34]. Moreover, CuFe2O4 as a highly dispersible catalyst improves the catalyst-BTA interaction and readily provides the photo-generated electrons to active PMS to yield SO4•− radicals [17].
To better understand the synergistic effect (SE) between the catalyst (CuFe2O4), UV, and PMS, the degradation efficiencies of binary and ternary processes were compared to individual processes as outlined in Table 2 [35]. A SE value higher than 1.0 elucidates the existence of a synergy between the components of the oxidation process. Table 2 shows that all the SE values for the binary and ternary processes are greater than 1.0, implying the existence of a synergistic effect in all the studied binary and ternary processes. These results confirm that combining UV, PMS, and CuFe2O4 provides an accelerated effect on the degradation of BTA in comparison to single and binary systems.

3.4. Influence of Operating Parameters on BTA Degradation

3.4.1. CCD Analysis

The CCD experiments and data on the experimental and predicted response values for BTA degradation as a function of catalyst dosage, PMS dosage, BTA initial concentration, and irradiation time are depicted in Table S1. Based on multiple regression analysis, the data showed quadratic polynomial prediction equation agreement in coded factor terms (±α, ±1, 0) as follows:
BTA removal (%) = 58.04 + 8x1 + 3.01x2 − 4.38x3 + 3.92x4 +
0.51x1x2 + 2.23x1x3 + 1.83x1x4 + 2.76x2x3 + 0.42x2x4
1.34x3x4 + 2.37x12 + 0.95x22 + 0.09x32 + 1.61x42
Equation (7) shows that the effects of catalyst loading (x1), PMS dosage (x2), and irradiation time (x4) on the predicted response were positive, whereas the effect of initial BTA concentration (x3) was negative. Additionally, the significant effect of x1, x2, and x4 on the BTA removal follows the order x1 > x4 > x2. ANOVA analysis (Table S2) shows that the quadratic regression model was highly significant (p-value > F-value: F-value = 24.18 and p-value = 0.0001) and reliable (Std. Dev. = 3.02). Furthermore, the R2 of the quadratic model was 0.9576, providing a 0.957 confidence level for response variability (Figure S1). It was further observed by the normality established between all residuals in the straight line (Figure S2), which can be randomly distributed in the range of ±3.00 (Figure S3), thus indicating an excellent approximation. In addition, R2adj of 0.918 was close to R2, verifying the goodness-of-fit of the model [36,37,38].
According to the prediction of CCD, the maximum removal of BTA was 81.46% (0.987 desirability), where the optimal values of influencing factors were catalyst loading (0.4 g L−1), PMS dosage (2 mM), BTA concentration (20 mg L−1), and irradiation time (70 min).

3.4.2. Interaction Analysis of Influential Factors

The effects of PMS dosage and catalyst loading on BTA degradation efficiency are shown in Figure 4A. It can be seen that higher initial PMS concentrations and catalyst loads promoted the removal rate of BTA within the specified ranges. This confirms the positive effect of both PMS and catalyst loading on the degradation efficiency. Such increased interaction between the catalyst and PMS results in the generation of a high yield of reactive radicals on the photocatalyst surface. Subsequently, the metal active species, i.e., Cu and Fe sites, decompose PMS to form the radical species, which contribute to the extensive oxidation of 30 mg L−1 of BTA within 50 min of irradiation time.
At the constant level of PMS and BTA concentration (Figure 4B), the response surface plot depicts that the catalyst load was not a limiting factor; higher catalyst loading under prolonged irradiation time caused the maximum removal of BTA. The BTA removal efficiency with the CuFe2O4/UV/PMS process significantly improved from 50.2% to 81.9% upon increasing the dosage of CuFe2O4 from 0.1 to 0.5 g L−1 after 50 min of reaction. This is because more SO4•− and OH radicals were generated via the activation of PMS on the CuFe2O4 surface (i.e., CuFe2O4 surf (e) and CuFe2O4 surf (h+)) and enhanced the adsorption process of BTA molecules onto the numerous available active sites. These results suggest that the efficacy was affected by the availability of the total surface area as well as the Cu/Fe active site interactions in the catalyst for reaction with PMS during irradiation.
Figure 4C shows the effect of BTA initial concentration and irradiation time on the system's performance. Obviously, the degradation decreased sharply with an increase in BTA concentration from 10 to 50 mg L−1. This is largely due to the fact that at constant catalyst loading, the amount of adsorption and catalytic sites for PMS activation and pollutant molecules remains the same despite the increasing amount of BTA molecules. Ultimately, this leads to a decline in the activation rate of PMS and the subsequent degradation of the BTA molecules. Moreover, as the BTA concentration increases, there is an emergence of fierce competition between parent BTA molecules and intermediates for the limited active sites and oxidative radicals on the catalyst surfaces. More specifically, increasing the concentration of BTA on the catalyst surface can decrease photocatalytic reactions that depend upon the direct contact of photo-induced electron/hole pairs to generate active species, which in turn hampers the degradation of BTA. Despite this, an increase in the degradation rate was observed upon extension of the irradiation time. For example, at an initial BTA concentration of 30 mg L−1, the degradation rates increased from 50.9% to 75.1% when the reaction time was increased from 10 to 90 min.

3.5. Recyclability Performance of the CuFe2O4/UV/PMS System

Recycling experiments were conducted to determine the stability and multiple reusability potential of the catalyst in the CuFe2O4/UV/PMS system. This is one of the strengths of the spinel-structured heterogeneous catalytic system. After each cycle, the used catalyst was recovered magnetically, followed by washing with deionized water three times and drying at 80 °C for 1 h. As shown in Figure 5A, the catalytic activity of CuFe2O4 towards BTA degradation could be maintained at as high as 79% after four photocatalytic cycles (280 min), accompanied by more than 40% TOC removal. Metal leaching detection indicated that CuFe2O4 also provided less Cu (0.33 mg L−1) and Fe (0.18 mg L−1) leaching (Figure 5B), corresponding to 0.10% and 0.06% of the total Cu and Fe contents in the catalyst, respectively. These findings reveal that the coordination activities of Cu/Fe pairs in the spinel structure effectively minimized the leaching of metal ions, which led to high stability and catalytic activity over several degradation cycles (less than 4% loss of activity). Despite the simplicity of the process, the CuFe2O4/UV/PMS system showed a good mineralization rate and removal efficiency of BTA compared to other CuFe2O4-based catalyst systems previously reported (Table 3).

3.6. Feasibility of the Process (Effect of Inorganic Anions)

An investigation of the interference of inorganic anions present in natural water and wastewater sources was carried out during BTA degradation to assess the practical applicability of the CuFe2O4/UV/PMS system. Therefore, BTA degradation was conducted in the presence of 10 mM of major anions, including bicarbonate (HCO3), chloride (Cl), sulfate (SO42−), and nitrate (NO3). Figure 6 reveals that the presence of the inorganic anions caused a suppressive effect on the BTA degradation. Generally, the effect of anions on removal efficiency can be described as follows: (i) catalyzing PMS through electron exchange reactions to generate some radicals with lower redox potential; (ii) competing with pollutant molecules for reactions with free radicals; and (iii) covering the reactive sites and absorbing photons on the CuFe2O4 surface, leading to the deactivation of reactive sites [32,41]. As shown in Figure 6, the process performance strongly decreased by 42.8% in the presence of HCO3 and 52.4% in the presence of Cl. However, SO42− and NO3 anions slightly reduced the process efficiency to 76.8% and 73.3%, respectively, which were almost close to the system efficiency without anions. As a result of low reaction kinetics with SO4•−/OH, the demoting roles of SO42− and NO3 as radical scavengers in BTA degradation were lower than those of Cl and HCO3, as shown by Equations (8)–(14). Moreover, through Equations (8) and (9) SO42− can be converted into the aqueous electron (e) and highly reactive SO4•−, respectively. In this way, an aqueous electron ( = −2.9 V) could attack halogenated organic compounds [21].
SO4•− + SO42− → S2O82− + e (4.4 × 108 M−1 s−1)
OH + SO42− → SO4•− + OH (3.5 × 105 M−1 s−1)
h+ + SO42− → SO4•−
SO4•− + e → SO42−
SO4•− + NO3 → NO3•− + SO42− (5.5 × 105 M−1 s−1)
OH + NO3 → NO3•− + OH (<5 × 105 M−1 s−1)
h+ + NO3 → NO3
According to Equations (15) and (16), chloride ions can be thermodynamically oxidized by SO4•− to chlorine radicals, and therefore chloride ions showed a significant radical scavenging role in the reaction. This decreases the process efficiency from 82.6 to 30.17% (Figure 6). Thus, the lower redox potential of Cl2•−/2Cl ( = 2.09 V) and Cl/Cl ( = 2.47 V) compared to that of SO4•−/SO42− ( = 2.5–3.1 V) may turn SO4•− into slower oxidants (i.e., Cl, ClOH•−, and Cl2•−). PMS ( = 1.75 V) could also react with chloride ions to form sulfate ions in Cl2 and HOCl ( Cl2/2Cl = 1.36 V and HOCl/Cl = 1.48 V) according to Equations (20) and (21) [42].
SO4•− + Cl ↔ SO42− + Cl (2.3 × 108 M−1 s−1)
OH + Cl ↔ ClOH•− (4.2 × 109 M−1 s−1)
ClOH•− + H+ → Cl + H2O (6.1 × 109 M−1 s−1)
h+ + Cl → Cl
Cl + Cl ↔ Cl2•− (7.8 × 109 M−1 s−1)
HSO5 + Cl ↔ SO42− + HOCl
HSO5 + 2Cl + H+ ↔ SO42− + Cl2 + H2O
As shown in Figure 6, a significant decrease (39.8%) in the degradation efficiency was observed in the presence of HCO3. This may be attributed to the fast scavenging of the photo-oxidizing species (SO4•−, OH, and h+) by HCO3, as well as the production of CO3•− via Equations (22)–(24), which mitigate the consumption of BTA [18]. Despite the high availability of HCO3•−/CO3•−, there was no enhancement effect on BTA removal because of their low oxidation capability [43]. Additionally, HCO3 was found to inhibit PMS activation.
SO4•− + HCO3 → SO42− + CO3•− + H+ (1.6 × 106 M−1 s−1)
OH + HCO3 → CO3•− + H2O (8.5 × 106 M−1 s−1)
h+ + HCO3 → HCO3•−

3.7. Radical Scavenging Experiments

Quenching tests were conducted to identify the predominant active species in the CuFe2O4/UV/PMS system responsible for BTA degradation (Figure 3). The quenching agents, including TBA, MeOH, BQ, and KI, were used in concentrations of 10 mM to trap OH, SO4•−, O2•−, and h+, respectively [9,16]. According to the different quenching reaction rates, MeOH can similarly quench both SO4•− and OH (k MeOH/•OH = 9.7 × 108 M−1 s−1, k MeOH/SO4•− = 3.2 × 106 M−1 s−1), while TBA, with 1000 times higher reactivity towards OH compared to SO4•− (k TBA/•OH = 3.8–7.6 × 108 M−1 s−1, k TBA/SO4•− = 4.0–9.1 × 105 M−1 s−1), preferentially scavenge OH radicals [44]. Figure 7 displays that the degradation efficiency decreased from 82.6% (without any quenchers) to 78 and 39.6% in the system with TBA and MeOH, respectively. Furthermore, it was found that the addition of KI resulted in a 65.4% reduction in process efficiency, while BQ led to a much stronger inhibitory effect (a 54.0% decrease), even more than TBA. These results confirmed the production of the reactive species OH, SO4•−, O2•−, and h+, and among them, SO4•− played the most predominant role during the BTA decomposition process in the CuFe2O4/UV/PMS system.

3.8. Mechanistic Discussion

Based on the active species capture experiments and some previous studies [9,13,22], a possible mechanism (Scheme 1) has been proposed to explain the charge transfer routes during the photocatalytic PMS activation leading to BTA degradation. The absorbed UV light caused the excitation of electrons in the valence band (VB) of the catalyst to the conduction band (CB) to produce photo-induced electron and hole pairs (e/h+) as depicted in Equation (25). Subsequently, the conduction band electrons (eCB) of CuFe2O4 could quickly reduce dissolved O2 to O2•− radicals (Equation (26)) based on the standard redox potential difference (E0eCB = −0.48 eV vs. NHE compared to E0O2/O2•− = −0.46 eV vs. NHE) [22]. At the same time, h+ could react with adsorbed H2O or OH to form OH (Equation (27)) [45]. O2•− could also trigger a series of redox reactions to produce OH (Equations (28) and (29)) [9].
CuFe2O4 + hv → hVB+ + eCB
CuFe2O4 (eCB) + O2 → O2•−
CuFe2O4 (hVB+) + H2O → OH + H+
O2•− + eCB + 2H+ → H2O2
H2O2 + eCBOH + OH
Additionally, it was observed that PMS further enhanced the catalytic reaction, which could be attributed to the chemical stability arising from the redox cycles of surface active centers. It was concluded that the process of CuFe2O4 catalyzed PMS for generating reactive species was initially related to the binding of hydroxyl groups (−OH) obtained from the dissociation of water on the surface metal (Fe or Cu) sites in the CuFe2O4 catalyst [27]. In this regard, HSO5 could form a bond with the surface Cu(II) of the catalyst via surface −OH displacement and generate a Cu(II)-(OH)OSO3 intermediate by the inner-sphere complexation (Equation 30). Therefore, the favorability of the electron transfer from Cu(II), with the obtained high electron density, to OH of HSO5 could lead directly to the production of SO4•− radical and a new surface −OH group that could bond to a higher valence Cu(III) ion (Equation (31)). Supposing that ≡Cu(III)−OH oxidized HSO5 to SO5•− bonded to copper initial valence (i.e., Cu(II)) (Equation (32)), then the combination of surface SO5•− moieties would again generate SO4•− via Equation (33). The efficient participation in the redox process could result in the reductant character for Fe(II), reducing ≡Cu(III) to ≡Cu(II) (Equation (34)), since E0≡Cu(III)/≡Cu(II) = 2.3 V is much higher than E0≡Fe(III)/≡Fe(II) = 0.77 V. The resulting ≡Fe(III) would be possibly turned to ≡Fe(II) during the process of HSO5 reduction (Equation 35), which would be subsequently oxidized to ≡Fe(III) by the production of SO4•− radicals (Equation (36)). Thus, redox mediators, i.e., Cu(II-III-II) and Fe(III-II-III), by maintaining the reaction cycles not only enhance the activity of CuFe2O4 via Equation (34) but also contribute to the further decomposition of PMS through Equations (35) and (36). From the hydroxylation of SO4•−,OH can be released according to Equations (37) and (38) [46]. Photocatalytic activation of PMS, under the action of photoinduced electrons, opened another route for the direct generation of OH and SO4•− radicals (Equation (39)). In this way, the photoinduced holes could be captured by PMS to produce SO4•− via subsequent self-reaction of SO5•− radicals (Equations (40) and (41)). On the other hand, during the reaction, photo-induced electrons under the different valence states of Cu–Fe could create a new equilibrium to obtain the cyclic Cu(III)/Cu(II) and Fe(III)/Fe(II) (Equation (42)). This process can lead to fast photocharge transfer in the presence of PMS and may explain the higher reactivity compared to the binary systems, namely CuFe2O4/UV and CuFe2O4/PMS (Figure 3). Thus, when combined with photocatalysis, PMS oxidation showed a synergistic effect by producing more free active species. A negligible amount of SO4•−, OH, O2•−, and h+ radicals that could have leached from the surface-bound layer to the solution bulk in the CuFe2O4/UV/PMS system has little effect on the mineralization of BTA (Equation (43)).
≡Cu(II)˗˗˗OH + HSO5 → ≡Cu(II)˗˗˗(OH)˗OSO3 + OH
≡Cu(II)˗˗˗(OH)˗OSO3 → ≡Cu(III)˗˗˗OH + SO4•−
≡Cu(III)˗˗˗OH + HSO5 → ≡Cu(II)˗˗˗SO5•− + H2O
2≡Cu(II)˗˗˗SO5•− + 2H2O → 2≡Cu(II)˗˗˗OH + O2 + 2SO4•− + 2H+
≡Cu(III)˗˗˗OH + ≡Fe(II) → ≡Cu(II)˗˗˗OH + ≡Fe(III)
≡Fe(III) + HSO5 → ≡Fe(II) + SO5•− + H+
≡Fe(II) + HSO5 → ≡Fe(III) + SO4•− + OH
SO4•− + H2O → OH + SO42− + H+
SO4•− + OHOH + SO42−
eCB + HSO5 → SO4•− + OH
hVB+ + HSO5 → SO5•− + H+
2SO5•− → 2SO4•− + O2
Cu(III)/Fe(III) + eCB → Cu(II)/Fe(II)
SO4•−/OH/O2•−/h+ + BTA → products → CO2 + H2O + NH3

3.9. Reaction Pathway of BTA Degradation

The BTA degradation pathway in the CuFe2O4/UV/PMS system was proposed and described in Figure 8 based on the literature [47,48,49]. As shown in Figure 8, radicals attack the active sites of the BTA molecule and may first cleave the triazole ring of the molecule at site N14. In this way, the nitrogen double bond was destroyed, and following the N–NH bond dissociation, it would form the yellow intermediate diazoimine. This primary diazo intermediate was a transient product, resulting in its stepwise oxidation to release a nitrogen molecule and produce the colorless biradical intermediate under irradiation. Based on the recombination reaction, aniline could be formed [5,47], which is in accordance with the identification of HPLC analysis using standard aniline. The structure and chemical formula of the identified intermediates are given in Table S3. In pathway (I), the intermediates aniline radicals (P1) and benzoquinonimine (P2) appeared. In this step, the aniline radicals could undergo direct polymerization to form dianiline (P3) leading to the opening of the aromatic ring of aniline to form maleic acid (P7) [48]. Alternatively, the aniline radicals could produce intermediate 4-aminophenol (P4), which is the main precursor for the formation of benzoquinone (P5) via the action of ROS. Additionally, benzoquinone (P5) can be formed through the hydrolysis reaction of benzoquinonimine (P2) that occurs concurrently with deamination. Moreover, nitrobenzene (P6) was observed through the catalytic activity of ROS on benzoquinonimine (P2), leading to the formation of maleic acid (P7) by further deamination, especially losing NO3 and finally being converted into CO2 and H2O [49]. Under the attack of the ROS, the deprotonation process on the external carbons could also occur to produce the ring-cleavage intermediate 3-aminoprop-2-en-1-ol (P8) (pathway II) [2].

4. Conclusions

Herein, photosynergistic activation of PMS with a heterogenous CuFe2O4 catalyst was assessed to study BTA degradation under UV light. The novel-designed system exhibited a 1.17- and 1.3-fold increase in catalytic activity for BTA degradation, reaching a high efficiency of 73%. This result was higher than that of CuFe2O4/PMS (62.2%) and CuFe2O4/UV (53.6%) binary systems, respectively, indicating a synergy effect between UV, PMS, and the catalyst. The radical quenching tests conducted in this study demonstrated the generation of multiple radicals, including OH, SO4•−, O2•−, and h+. These radicals were generated through the conversion of Cu(II)/Cu(III) and Fe(III)/Fe(II) on the surface of CuFe2O4 with SO4•− playing a dominant role in the radical generation process. This resulted in negligible leaching of metals and the excellent recyclability of synthesized catalysts. The samples prepared in this study demonstrated excellent catalytic performance and low levels of Fe and Cu leaching over the course of four consecutive runs. Collectively, this study presented a feasible approach for improving the catalytic performance of wastewater treatment systems for environmental remediation.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/toxics11050429/s1. Table S1: CCD experiments and observed and predicted removal efficiencies using CuFe2O4/UV/PMS system; Table S2: Results of ANOVA for response surface quadratic model; Figure S1: Correlation between predicted and experimental values of BTA degradation variability; Figure S2: The normal probability plot of the internally studentized residuals; Figure S3: The experimental run number versus studentized residual data; Table S3 The structure and chemical formula of detected intermediates from BTA degradation in the CuFe2O4/UV/PMS system.

Author Contributions

Conceptualization, M.G. and B.K.; methodology, B.K.; software, B.K.; validation, M.G., B.K. and N.T.; formal analysis, M.G.; investigation, B.K.; resources, M.G.; data curation, G.M.; writing—original draft preparation, M.G. and B.K.; writing—review and editing, G.M. and N.T.; visualization, N.T.; supervision, B.K.; project administration, M.G.; funding acquisition, M.G. All authors have read and agreed to the published version of the manuscript.

Funding

This work is financed by Zabol University of Medical Sciences (IR.ZBMU.REC.1399.085).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data will be made available on request.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Bahnmueller, S.; Loi, C.H.; Linge, K.L.; Von Gunten, U.; Canonica, S. Degradation rates of benzotriazoles and benzothiazoles under UV-C irradiation and the advanced oxidation process UV/H2O2. Water Res. 2015, 74, 143–154. [Google Scholar] [CrossRef] [PubMed]
  2. Xu, J.; Li, L.; Guo, C.; Zhang, Y.; Wang, S. Removal of benzotriazole from solution by BiOBr photocatalysis under simulated solar irradiation. Chem. Eng. J. 2013, 221, 230–237. [Google Scholar] [CrossRef]
  3. Ye, J.; Zhou, P.; Chen, Y.; Ou, H.; Liu, J.; Li, C.; Li, Q. Degradation of 1H-benzotriazole using ultraviolet activating persulfate: Mechanisms, products and toxicological analysis. Chem. Eng. J. 2018, 334, 1493–1501. [Google Scholar] [CrossRef]
  4. Felis, E.; Sochacki, A.; Magiera, S. Degradation of benzotriazole and benzothiazole in treatment wetlands and by artificial sunlight. Water Res. 2016, 104, 441–448. [Google Scholar] [CrossRef]
  5. Ding, Y.; Yang, C.; Zhu, L.; Zhang, J. Photoelectrochemical activity of liquid phase deposited TiO2 film for degradation of benzotriazole. J. Hazard. Mater. 2010, 175, 96–103. [Google Scholar] [CrossRef]
  6. Huang, H.; Wang, H.-L.; Shi, S.-B.; Jiang, W.-F. In-situ fabrication of AgI/AgnMoxO3x+n/2/g-C3N4 ternary composite photocatalysts for benzotriazole degradation: Tuning the heterostructure, photocatalytic activity and photostability by the degree of molybdate polymerization. Sep. Purif. Technol. 2023, 307, 122874. [Google Scholar] [CrossRef]
  7. Ahmadi, M.; Kakavandi, B.; Jorfi, S.; Azizi, M. Oxidative degradation of aniline and benzotriazole over PAC@ FeIIFe2IIIO4: A recyclable catalyst in a heterogeneous photo-Fenton-like system. J. Photochem. Photobiol. A Chem. 2016, 336, 42–53. [Google Scholar] [CrossRef]
  8. Ghanbari, F.; Khatebasreh, M.; Mahdavianpour, M.; Lin, K.-Y.A. Oxidative removal of benzotriazole using peroxymonosulfate/ozone/ultrasound: Synergy, optimization, degradation intermediates and utilizing for real wastewater. Chemosphere 2020, 244, 125326. [Google Scholar] [CrossRef]
  9. Sun, Q.; Wang, X.; Liu, Y.; Xia, S.; Zhao, J. Activation of peroxymonosulfate by a floating oxygen vacancies-CuFe2O4 photocatalyst under visible light for efficient degradation of sulfamethazine. Sci. Total Environ. 2022, 824, 153630. [Google Scholar] [CrossRef]
  10. Wei, Y.; Liu, H.; Liu, C.; Luo, S.; Liu, Y.; Yu, X.; Ma, J.; Yin, K.; Feng, H. Fast and efficient removal of As (III) from water by CuFe2O4 with peroxymonosulfate: Effects of oxidation and adsorption. Water Res. 2019, 150, 182–190. [Google Scholar] [CrossRef]
  11. Ding, Y.; Tang, H.; Zhang, S.; Wang, S.; Tang, H. Efficient degradation of carbamazepine by easily recyclable microscaled CuFeO2 mediated heterogeneous activation of peroxymonosulfate. J. Hazard. Mater. 2016, 317, 686–694. [Google Scholar] [CrossRef]
  12. Lutze, H.V.; Kerlin, N.; Schmidt, T.C. Sulfate radical-based water treatment in presence of chloride: Formation of chlorate, inter-conversion of sulfate radicals into hydroxyl radicals and influence of bicarbonate. Water Res. 2015, 72, 349–360. [Google Scholar] [CrossRef]
  13. Li, X.; Chen, T.; Qiu, Y.; Zhu, Z.; Zhang, H.; Yin, D. Magnetic dual Z-scheme g-C3N4/BiVO4/CuFe2O4 heterojunction as an efficient visible-light-driven peroxymonosulfate activator for levofloxacin degradation. Chem. Eng. J. 2023, 452, 139659. [Google Scholar] [CrossRef]
  14. Wang, Z.; Bush, R.T.; Sullivan, L.A.; Chen, C.; Liu, J. Selective oxidation of arsenite by peroxymonosulfate with high utilization efficiency of oxidant. Environ. Sci. Technol. 2014, 48, 3978–3985. [Google Scholar] [CrossRef]
  15. Anipsitakis, G.P.; Dionysiou, D.D. Degradation of organic contaminants in water with sulfate radicals generated by the conjunction of peroxymonosulfate with cobalt. Environ. Sci. Technol. 2003, 37, 4790–4797. [Google Scholar] [CrossRef]
  16. Wu, X.; Sun, D.; Ma, H.; Ma, C.; Zhang, X.; Hao, J. Activation of peroxymonosulfate by magnetic CuFe2O4@ ZIF-67 composite catalyst for the study on the degradation of methylene blue. Colloids Surf. A Physicochem. Eng. Asp. 2022, 637, 128278. [Google Scholar] [CrossRef]
  17. Dong, X.; Ren, B.; Sun, Z.; Li, C.; Zhang, X.; Kong, M.; Zheng, S.; Dionysiou, D.D. Monodispersed CuFe2O4 nanoparticles anchored on natural kaolinite as highly efficient peroxymonosulfate catalyst for bisphenol A degradation. Appl. Catal. B Environ. 2019, 253, 206–217. [Google Scholar] [CrossRef]
  18. Fayyaz, A.; Saravanakumar, K.; Talukdar, K.; Kim, Y.; Yoon, Y.; Park, C.M. Catalytic oxidation of naproxen in cobalt spinel ferrite decorated Ti3C2Tx MXene activated persulfate system: Mechanisms and pathways. Chem. Eng. J. 2021, 407, 127842. [Google Scholar] [CrossRef]
  19. Abdelbasir, S.M.; Elseman, A.M.; Harraz, F.A.; Ahmed, Y.M.; El-Sheikh, S.M.; Rashad, M.M. Superior UV-light photocatalysts of nano-crystalline (Ni or Co) FeWO4: Structure, optical characterization and synthesis by a microemulsion method. New J. Chem. 2021, 45, 3150–3159. [Google Scholar] [CrossRef]
  20. Yang, S.; Qiu, X.; Jin, P.; Dzakpasu, M.; Wang, X.C.; Zhang, Q.; Yang, L.; Ding, D.; Wang, W.; Wu, K. MOF-templated synthesis of CoFe2O4 nanocrystals and its coupling with peroxymonosulfate for degradation of bisphenol A. Chem. Eng. J. 2018, 353, 329–339. [Google Scholar] [CrossRef]
  21. Li, Z.; Wang, F.; Zhang, Y.; Lai, Y.; Fang, Q.; Duan, Y. Activation of peroxymonosulfate by CuFe2O4-CoFe2O4 composite catalyst for efficient bisphenol a degradation: Synthesis, catalytic mechanism and products toxicity assessment. Chem. Eng. J. 2021, 423, 130093. [Google Scholar] [CrossRef]
  22. Xin, J.; Liu, Y.; Niu, L.; Zhang, F.; Li, X.; Shao, C.; Li, X.; Liu, Y. Three-dimensional porous CuFe2O4 for visible-light-driven peroxymonosulfate activation with superior performance for the degradation of tetracycline hydrochloride. Chem. Eng. J. 2022, 445, 136616. [Google Scholar] [CrossRef]
  23. Ren, Y.; Lin, L.; Ma, J.; Yang, J.; Feng, J.; Fan, Z. Sulfate radicals induced from peroxymonosulfate by magnetic ferrospinel MFe2O4 (M = Co, Cu, Mn, and Zn) as heterogeneous catalysts in the water. Appl. Catal. B Environ. 2015, 165, 572–578. [Google Scholar] [CrossRef]
  24. He, S.; Chen, Y.; Li, X.; Zeng, L.; Zhu, M. Heterogeneous photocatalytic activation of persulfate for the removal of organic contaminants in water: A critical review. ACS EST Eng. 2022, 2, 527–546. [Google Scholar] [CrossRef]
  25. Hasija, V.; Nguyen, V.-H.; Kumar, A.; Raizada, P.; Krishnan, V.; Khan, A.A.P.; Singh, P.; Lichtfouse, E.; Wang, C.; Huong, P.T. Advanced activation of persulfate by polymeric g-C3N4 based photocatalysts for environmental remediation: A review. J. Hazard. Mater. 2021, 413, 125324. [Google Scholar] [CrossRef]
  26. Kakavandi, B.; Alavi, S.; Ghanbari, F.; Ahmadi, M. Bisphenol A degradation by peroxymonosulfate photo-activation coupled with carbon-based cobalt ferrite nanocomposite: Performance, upgrading synergy and mechanistic pathway. Chemosphere 2022, 287, 132024. [Google Scholar] [CrossRef]
  27. Zhang, T.; Zhu, H.; Croué, J.-P. Production of sulfate radical from peroxymonosulfate induced by a magnetically separable CuFe2O4 spinel in water: Efficiency, stability, and mechanism. Environ. Sci. Technol. 2013, 47, 2784–2791. [Google Scholar] [CrossRef]
  28. Wang, Y.; Tian, D.; Chu, W.; Li, M.; Lu, X. Nanoscaled magnetic CuFe2O4 as an activator of peroxymonosulfate for the degradation of antibiotics norfloxacin. Sep. Purif. Technol. 2019, 212, 536–544. [Google Scholar] [CrossRef]
  29. Ding, Y.; Zhu, L.; Wang, N.; Tang, H. Sulfate radicals induced degradation of tetrabromobisphenol A with nanoscaled magnetic CuFe2O4 as a heterogeneous catalyst of peroxymonosulfate. Appl. Catal. B Environ. 2013, 129, 153–162. [Google Scholar] [CrossRef]
  30. Guan, Y.-H.; Ma, J.; Ren, Y.-M.; Liu, Y.-L.; Xiao, J.-Y.; Lin, L.-q.; Zhang, C. Efficient degradation of atrazine by magnetic porous copper ferrite catalyzed peroxymonosulfate oxidation via the formation of hydroxyl and sulfate radicals. Water Res. 2013, 47, 5431–5438. [Google Scholar] [CrossRef]
  31. Golshan, M.; Kakavandi, B.; Ahmadi, M.; Azizi, M. Photocatalytic activation of peroxymonosulfate by TiO2 anchored on cupper ferrite (TiO2@ CuFe2O4) into 2,4-D degradation: Process feasibility, mechanism and pathway. J. Hazard. Mater. 2018, 359, 325–337. [Google Scholar] [CrossRef]
  32. Lei, X.; You, M.; Pan, F.; Liu, M.; Yang, P.; Xia, D.; Li, Q.; Wang, Y.; Fu, J. CuFe2O4@ GO nanocomposite as an effective and recoverable catalyst of peroxymonosulfate activation for degradation of aqueous dye pollutants. Chin. Chem. Lett. 2019, 30, 2216–2220. [Google Scholar] [CrossRef]
  33. Xu, Y.; Ai, J.; Zhang, H. The mechanism of degradation of bisphenol A using the magnetically separable CuFe2O4/peroxymonosulfate heterogeneous oxidation process. J. Hazard. Mater. 2016, 309, 87–96. [Google Scholar] [CrossRef]
  34. Wang, Y.; Sun, H.; Ang, H.M.; Tadé, M.O.; Wang, S. Facile synthesis of hierarchically structured magnetic MnO2/ZnFe2O4 hybrid materials and their performance in heterogeneous activation of peroxymonosulfate. ACS Appl. Mater. Interfaces 2014, 6, 19914–19923. [Google Scholar] [CrossRef]
  35. Tian, N.; Giannakis, S.; Akbarzadeh, L.; Hasanvandian, F.; Dehghanifard, E.; Kakavandi, B. Improved catalytic performance of ZnO via coupling with CoFe2O4 and carbon nanotubes: A new, photocatalysis-mediated peroxymonosulfate activation system, applied towards Cefixime degradation. J. Environ. Manag. 2023, 329, 117022. [Google Scholar] [CrossRef]
  36. Liu, F.; Wang, X.; Liu, Z.; Miao, F.; Xu, Y.; Zhang, H. Peroxymonosulfate enhanced photocatalytic degradation of Reactive Black 5 by ZnO-GAC: Key influencing factors, stability and response surface approach. Sep. Purif. Technol. 2021, 279, 119754. [Google Scholar] [CrossRef]
  37. Liu, X.; Zhou, J.; Liu, D.; Li, L.; Liu, W.; Liu, S.; Feng, C. Construction of Z-scheme CuFe2O4/MnO2 photocatalyst and activating peroxymonosulfate for phenol degradation: Synergistic effect, degradation pathways, and mechanism. Environ. Res. 2021, 200, 111736. [Google Scholar] [CrossRef]
  38. Kakavandi, B.; Ahmadi, M. Efficient treatment of saline recalcitrant petrochemical wastewater using heterogeneous UV-assisted sono-Fenton process. Ultrason. SonoChem. 2019, 56, 25–36. [Google Scholar] [CrossRef]
  39. Jia, Y.; Yang, K.; Zhang, Z.; Gu, P.; Liu, S.; Li, M.; Wang, X.; Yin, Y.; Zhang, Z.; Wang, T. Heterogeneous activation of peroxymonosulfate by magnetic hybrid CuFe2O4@ N-rGO for excellent sulfamethoxazole degradation: Interaction of CuFe2O4 with N-rGO and synergistic catalytic mechanism. Chemosphere 2023, 313, 137392. [Google Scholar] [CrossRef]
  40. Liu, X.; Zhou, J.; Wang, G.; Liu, D.; Liu, S. Construction of Cu-Fe bimetallic oxide/biochar/Ag3PO4 heterojunction for improving photocorrosion resistance and photocatalytic performance achieves efficient removal of phenol. Appl. Surf. Sci. 2022, 592, 153307. [Google Scholar] [CrossRef]
  41. Li, D.; Yao, Z.; Lin, J.; Tian, W.; Zhang, H.; Duan, X.; Wang, S. Nano-sized FeVO4 1.1H2O and FeVO4 for peroxymonosulfate activation towards enhanced photocatalytic activity. J. Environ. Chem. Eng. 2022, 10, 107199. [Google Scholar] [CrossRef]
  42. Anipsitakis, G.P.; Dionysiou, D.D.; Gonzalez, M.A. Cobalt-mediated activation of peroxymonosulfate and sulfate radical attack on phenolic compounds. Implications of chloride ions. Environ. Sci. Technol. 2006, 40, 1000–1007. [Google Scholar] [CrossRef] [PubMed]
  43. Nie, M.; Yang, Y.; Zhang, Z.; Yan, C.; Wang, X.; Li, H.; Dong, W. Degradation of chloramphenicol by thermally activated persulfate in aqueous solution. Chem. Eng. J. 2014, 246, 373–382. [Google Scholar] [CrossRef]
  44. Li, Z.; Tang, X.; Huang, G.; Luo, X.; He, D.; Peng, Q.; Huang, J.; Ao, M.; Liu, K. Bismuth MOFs based hierarchical Co3O4-Bi2O3 composite: An efficient heterogeneous peroxymonosulfate activator for azo dyes degradation. Sep. Purif. Technol. 2020, 242, 116825. [Google Scholar] [CrossRef]
  45. Golshan, M.; Zare, M.; Goudarzi, G.; Abtahi, M.; Babaei, A.A. Fe3O4@ HAP-enhanced photocatalytic degradation of Acid Red73 in aqueous suspension: Optimization, kinetic, and mechanism studies. Mater. Res. Bull. 2017, 91, 59–67. [Google Scholar] [CrossRef]
  46. Li, C.; Huang, Y.; Dong, X.; Sun, Z.; Duan, X.; Ren, B.; Zheng, S.; Dionysiou, D.D. Highly efficient activation of peroxymonosulfate by natural negatively-charged kaolinite with abundant hydroxyl groups for the degradation of atrazine. Appl. Catal. B Environ. 2019, 247, 10–23. [Google Scholar] [CrossRef]
  47. Serdechnova, M.; Ivanov, V.L.; Domingues, M.R.M.; Evtuguin, D.V.; Ferreira, M.G.; Zheludkevich, M.L. Photodegradation of 2-mercaptobenzothiazole and 1, 2, 3-benzotriazole corrosion inhibitors in aqueous solutions and organic solvents. Phys. Chem. Chem. Phys. 2014, 16, 25152–25160. [Google Scholar] [CrossRef]
  48. Li, Y.; Wang, F.; Zhou, G.; Ni, Y. Aniline degradation by electrocatalytic oxidation. Chemosphere 2003, 53, 1229–1234. [Google Scholar] [CrossRef]
  49. Jorfi, S.; Kakavandi, B.; Motlagh, H.R.; Ahmadi, M.; Jaafarzadeh, N. A novel combination of oxidative degradation for benzotriazole removal using TiO2 loaded on FeIIFe2IIIO4@C as an efficient activator of peroxymonosulfate. Appl. Catal. B Environ. 2017, 219, 216–230. [Google Scholar] [CrossRef]
Figure 1. FESEM image (A) and EDS spectrum (B) of CuFe2O4; TEM micrographs of CuFe2O4 (C,D) with different scale bars.
Figure 1. FESEM image (A) and EDS spectrum (B) of CuFe2O4; TEM micrographs of CuFe2O4 (C,D) with different scale bars.
Toxics 11 00429 g001
Figure 2. (A) XRD patterns, (B) N2 adsorption-desorption isotherms (insert: pore size distribution), (C,D) magnetic hysteresis loops, and (E) UV−Vis absorption spectra of CuFe2O4. In (B), it is possible to include the reference diffractogram for CuFe2O4 for easy comparison.
Figure 2. (A) XRD patterns, (B) N2 adsorption-desorption isotherms (insert: pore size distribution), (C,D) magnetic hysteresis loops, and (E) UV−Vis absorption spectra of CuFe2O4. In (B), it is possible to include the reference diffractogram for CuFe2O4 for easy comparison.
Toxics 11 00429 g002
Figure 3. Removal efficiencies of BTA in different reaction systems. Conditions: BTA = 40 mg L−1, catalyst (CuFe2O4) = 0.2 g L−1 and PMS = 1.5 mM during the 60 min reaction time.
Figure 3. Removal efficiencies of BTA in different reaction systems. Conditions: BTA = 40 mg L−1, catalyst (CuFe2O4) = 0.2 g L−1 and PMS = 1.5 mM during the 60 min reaction time.
Toxics 11 00429 g003
Figure 4. Effect of operating factors on the BTA removal rate over the CuFe2O4/UV/PMS system: catalyst dosage vs. PMS concentration (A), catalyst dosage vs. reaction time (B), and BTA concentration vs. reaction time (C). Other variables were kept at zero level.
Figure 4. Effect of operating factors on the BTA removal rate over the CuFe2O4/UV/PMS system: catalyst dosage vs. PMS concentration (A), catalyst dosage vs. reaction time (B), and BTA concentration vs. reaction time (C). Other variables were kept at zero level.
Toxics 11 00429 g004
Figure 5. Performance of the recycled catalysts in the CuFe2O4/UV/PMS system on BTA mineralization (A) and Cu and Fe leaching from CuFe2O4 after each cycle (B). Conditions: C0 = 20 mg L−1, CuFe2O4 = 0.4 g L−1, PMS = 2 mM, and time = 70 min.
Figure 5. Performance of the recycled catalysts in the CuFe2O4/UV/PMS system on BTA mineralization (A) and Cu and Fe leaching from CuFe2O4 after each cycle (B). Conditions: C0 = 20 mg L−1, CuFe2O4 = 0.4 g L−1, PMS = 2 mM, and time = 70 min.
Toxics 11 00429 g005
Figure 6. Percentage of BTA removal in the presence of different anions. Conditions: C0 = 20 mg L−1, CuFe2O4 = 0.4 g L−1, PMS = 2 mM, time = 70 min.
Figure 6. Percentage of BTA removal in the presence of different anions. Conditions: C0 = 20 mg L−1, CuFe2O4 = 0.4 g L−1, PMS = 2 mM, time = 70 min.
Toxics 11 00429 g006
Figure 7. The effects of different radical scavengers on BTA degradation. Conditions: C0 = 20 mg L−1, CuFe2O4 = 0.4 g L−1, PMS = 2 mM, time = 70 min.
Figure 7. The effects of different radical scavengers on BTA degradation. Conditions: C0 = 20 mg L−1, CuFe2O4 = 0.4 g L−1, PMS = 2 mM, time = 70 min.
Toxics 11 00429 g007
Scheme 1. Schematic diagram of the reaction mechanism for degradation of BTA by the CuF2O4/UV/PMS process.
Scheme 1. Schematic diagram of the reaction mechanism for degradation of BTA by the CuF2O4/UV/PMS process.
Toxics 11 00429 sch001
Figure 8. Schematic degradation pathways of BTA during the CuFe2O4/UV/PMS process based on the potential consequences.
Figure 8. Schematic degradation pathways of BTA during the CuFe2O4/UV/PMS process based on the potential consequences.
Toxics 11 00429 g008
Table 1. Four independent factors and their ranges at various levels.
Table 1. Four independent factors and their ranges at various levels.
Independent FactorsUnitSymbolsRanges and Levels
−αLow (−1)Middle (0)High (+1)
Catalyst loadingg L−1x10.10.20.30.40.5
PMS dosagemMx20.511.522.5
Initial BTA
concentration
mg L−1x31020304050
Irradiation timeminx41030507090
Table 2. Summary of the synergistic effect of different systems towards BTA elimination.
Table 2. Summary of the synergistic effect of different systems towards BTA elimination.
No.SE EquationSE Value
1 S E = k P M S / U V k P M S + k U V 1.68
2 S E = k C u F e 2 O 4 / U V k C u F e 2 O 4 + k U V 1.38
3 S E = k C u F e 2 O 4 / P M S k C u F e 2 O 4 + k P M S 1.51
4 S E = k C u F e 2 O 4 / U V / P M S k C u F e 2 O 4 + k P M S + k U V 1.50
5 S E = k C u F e 2 O 4 / U V / P M S k C u F e 2 O 4 / U V + k P M S 1.15
6 S E = k C u F e 2 O 4 / U V / P M S k C u F F e 2 O 4 / P M S + k U V 1.04
Table 3. The degradation of pollutants in other similar oxidation systems with various catalysts.
Table 3. The degradation of pollutants in other similar oxidation systems with various catalysts.
SystemPollutant (mg/L)Catalyst (g/L)Oxidant (mM)Removal Efficiency (%)Mineralization Rate (%)Metal Ion Leaching (mg/L)Ref.
CuFe@NG/PMSSulfamethoxazole (10)0.30.493.15 in 60 min31.96[Cu] = 0.25, [Fe] = -[39]
CuFe2O4@GO/PMSMethylene blue (20)0.20.893.3 in 30 min-[Cu] = 0.3, [Fe] = 0.3[32]
OVs-CFEp/Vis/PMSSulfamethazine (10)10.395 in 90 min56[Cu] < 0.1, [Fe] < 0.1[9]
CuFe2O4 3DPs/Vis/PMSTetracycline hydrochloride (20)0.250.293 in 30 min [Cu] = 0.2, [Fe] = 0.2[22]
CuFe2O4-CoFe2O4/PMSBPA (20)0.2198.7 in 110 min72.5 [21]
CuFe2O4/kaolinite/PMSBPA (50)0.50.597 in 60 min55[Cu] = 0.27, [Fe] = 0.01[17]
CuFe2O4/Biochar/Ag3PO4/VisPhenol (20)0.15-100 in 18 min [40]
CN/BVO/CFO/Vis/PMSLevofloxacin (10)0.2196.2 in 60 min67[Cu] = 0.08, [Fe] = 0.02[13]
CuFe2O4/MnO2/Vis/PMSPhenol (100)0.50.5100 in 30 min62.2[Cu] = 0.026, [Fe] = 0.012[37]
CuFe2O4/UV/PMSBTA (20)0.4282.3 in 70 min53.2[Cu] = 0.33, [Fe] 0.18This study
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

Golshan, M.; Tian, N.; Mamba, G.; Kakavandi, B. Synergetic Photocatalytic Peroxymonosulfate Oxidation of Benzotriazole by Copper Ferrite Spinel: Factors and Mechanism Analysis. Toxics 2023, 11, 429. https://doi.org/10.3390/toxics11050429

AMA Style

Golshan M, Tian N, Mamba G, Kakavandi B. Synergetic Photocatalytic Peroxymonosulfate Oxidation of Benzotriazole by Copper Ferrite Spinel: Factors and Mechanism Analysis. Toxics. 2023; 11(5):429. https://doi.org/10.3390/toxics11050429

Chicago/Turabian Style

Golshan, Masoumeh, Na Tian, Gcina Mamba, and Babak Kakavandi. 2023. "Synergetic Photocatalytic Peroxymonosulfate Oxidation of Benzotriazole by Copper Ferrite Spinel: Factors and Mechanism Analysis" Toxics 11, no. 5: 429. https://doi.org/10.3390/toxics11050429

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