Previous Article in Journal
Plasticized Polylactide Film Coating Formation from Redispersible Particles
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Chromium Ferrite Supported on Activated Carbon from Olive Mill Solid Waste for the Photo-Fenton Degradation of Pollutants from Wastewater Using LED Irradiation

1
Laboratory of Applied Studies for Sustainable Development and Renewable Energy (LEADDER), Faculty of Sciences, Doctoral School for Science and Technology (EDST), Lebanese University, Hariri Campus, Hadath P.O. Box 90656, Lebanon
2
Laboratory of Materials, Catalysis, Environment, and Analytical Methods (MCEMA), Faculty of Sciences, Lebanese University, Hadath P.O. Box 6573, Lebanon
3
Bahaa and Walid Bassatne Department of Chemical Engineering and Advanced Energy, Faculty of Engineering and Architecture, American University of Beirut, Beirut P.O. Box 11-0236, Lebanon
4
College of Engineering and Technology, American University of the Middle East, Egaila 54200, Kuwait
5
Faculty of Science and Engineering, Maastricht University, P.O. Box 616, 6200 MD Maastricht, The Netherlands
6
Institut de Science des Matériaux de Mulhouse, Université de Haute-Alsace, CNRS, IS2M UMR 7361, F-68100 Mulhouse, France
*
Author to whom correspondence should be addressed.
AppliedChem 2025, 5(3), 15; https://doi.org/10.3390/appliedchem5030015
Submission received: 2 June 2025 / Revised: 28 June 2025 / Accepted: 7 July 2025 / Published: 11 July 2025

Abstract

In this study, chromium ferrite (FeCr; CrFe2O4) nanoparticles supported on activated carbon (AC), obtained from agricultural olive mill solid waste, were synthesized via a simple hydrothermal process. The structural, morphological, optical, and chemical properties of the FeCr/AC composite were characterized using XRD, SEM, EDX, DRS, BET, and FTIR techniques. The FeCr/AC composite was applied as a heterogeneous photo-Fenton catalyst for the degradation of methylene blue (MB) dye in an aqueous solution under 25 W visible-light LED irradiation. Critical operational factors, such as FeCr/AC dosage, pH, MB concentration, and H2O2 levels, were optimized. Under optimal conditions, 97.56% of MB was removed within 120 min of visible-light exposure, following pseudo-first-order kinetics. The composite also exhibited high efficiency in degrading methyl orange dye (95%) and tetracycline antibiotic (88%) within 180 min, with corresponding first-order rate constants of 0.0225 min−1 and 0.0115 min−1, respectively. This study highlights the potential of FeCr/AC for treating water contaminated with dyes and pharmaceuticals, in line with the Sustainable Development Goals (SDGs) for water purification.

Graphical Abstract

1. Introduction

The accumulation of hazardous and non-degradable organic compounds in wastewater, as a result of the growing population and increased industrialization, has been recognized as a serious concern in recent years. By 2030, approximately 47% of the global population will face the challenge of clean freshwater scarcity [1]. Global water resources are being contaminated by a variety of contaminants, such as heavy metals, colorants, detergents, drugs, pesticides, and phenols, discharged by industries and municipalities [2]. Among these, dye effluents have garnered significant researchers’ attention due to their complex structure, high stability, and resistance to conventional treatment methods. Organic dyes have been extensively employed in various industrial activities, including printing, plastics, painting, textiles, and food [3]. According to the statistical findings, around 30,000 tons of industrial dyes are released into the ecosystem annually [4,5]. However, many types of dyes are lethal, carcinogenic, and resistant to degradation, resulting in severe environmental pollution and posing irreparable damage to human health and the ecosystem [6,7]. Recently, the pursuit of efficient techniques to reduce organic pollutants in wastewater has prompted the investigation of numerous innovative technologies. Over the past decades, conventional methods such as coagulation, electrocoagulation, adsorption, filtration, bioremediation, electrolysis, and chemical oxidation have been the primary techniques for treating large-scale dye-contaminated wastewater [8,9,10,11,12,13]. The adsorption process is a fast, facile, and economic method for color removal from aquatic bodies [14,15,16]. However, such a method has a major limitation of secondary waste being produced after the contaminants are adsorbed. Additionally, there are environmental concerns related to the reuse of adsorbents for successive batches, the leaching or desorption of pollutants, and the disposal of spent adsorbents. Advanced oxidation processes (AOPs) indeed stand out as a potent method for addressing these challenges [17]. AOPs are used to break down many toxic and biologically persistent organic pollutants in water to acceptable levels, without generating harmful by-products [18]. This method relies on the production of •OH having an oxidation potential of 2.8 V, for the mineralization of pollutants into CO2, H2O, and inorganic ions [19,20]. Some examples of AOPs include, photocatalysis, Fenton, photo-Fenton, sono-photo-Fenton, and ozonation [21,22,23]. The heterogeneous photo-Fenton process is considered an economical and environmentally friendly technique for breaking down resistant contaminants due to its fast reaction rate, broad pH range application, and excellent recyclability [24,25,26,27,28].
Currently, spinel ferrite nanoparticles with the general molecular formula MFe2O4 (M = Ni, Mg, Zn, Co, Cu, Au) are widely valued due to their unique characteristics [29,30,31,32,33,34]. These include low price, good chemical and thermal stability, strong electrical resistivity, and magnetic anisotropy [35]. In particular, MFe2O4 has been widely employed in diverse applications, such as drug delivery, magnetic resonance image (MRI), cancer diagnosis, electronic devices, adsorption, and photocatalysis [36,37]. As one of its advantages, MFe2O4 semiconductors exhibit light-response properties, demonstrating efficient photocatalytic activity comparable to the most commonly used semiconductors (TiO2, CdS, ZnO, g-C3N4, and SnO2). Additionally, MFe2O4 has a lower band gap (2–3 eV), which permits them to absorb light in the visible region [38]. Ferrite nanoparticles still encounter several challenges as they tend to accumulate as a result of Van der Waals forces and magnetic dipolar interactions [39,40]. Also, the leaching of metals in water and the fast electron–hole recombination restrict their application in photocatalytic applications. Thus, the aforementioned issues can be mitigated by loading metal ferrites into suitable support materials.
Activated carbon (AC) is commonly employed for water purification either as an adsorbent or a carrier [41,42]. Its high surface area and porosity allow AC to effectively, quickly, and completely capture targeted contaminants, offering a notable advantage over other support systems playing this role [43]. The sustained high cost of commercial AC limits its application at a large industrial scale [44]. To address this limitation, lignocellulosic biomass has emerged as a sustainable, renewable, and cost-effective alternative for the development of carbon-based materials. Olive mill, agricultural solid waste (OMSW), a by-product of olive oil production, is a lignocellulosic biomass rich in cellulose, hemicellulose, and lignin [45]. It can be valorized into low-cost AC as one of the eco-friendly methods used to reduce their negative environmental impacts. Around 20 million tons of OMSW are produced every year [46]. In photocatalysis application, AC is extensively used as a support material for the deposition of metals and metal oxides. The recent research has shown that the combination of AC with suitable semiconductors improves the efficacy in catalytic oxidation and acts as a highly stable photocatalyst for decomposing numerous contaminants [47,48]. Additionally, the loading of semiconductors on AC resulted in a band gap reduction, increased solar energy collection, improved pollutant adsorption, and decreased electron–hole recombination [49,50]. Hamieh et al. [51] in their study developed a hybrid mixture of FeCr-SBA-15/AC to investigate the combined effect of adsorption and photo-Fenton processes on the degradation of methyl orange (MO) dye. The AC in the mixture improved MO adsorption and reduced the electron–hole recombination, thereby improving the overall photocatalytic activity. Therefore, it is feasible to enhance the photo-Fenton activity of metal ferrites by coupling them with AC.
In the present work, a composite mixture of FeCr/AC was prepared by the hydrothermal method. The objective of this research was to investigate the photocatalytic performance of the hybrid system while assessing the stability of the CrFe2O4 phase through leaching measurements under reaction conditions, with the aim of exploring potential improvements in the band gap and operational stability compared to conventional systems. Additionally, the presence of FeCr facilitates charge separation and enhances hydroxyl radical generation, which contributes to an improved photocatalytic performance [52]. On the contrary, the high surface area of AC plays a crucial role in improving catalytic activity. The production of AC with desirable physicochemical properties from agricultural biomass waste aligns with the ongoing efforts in sustainable development and the promotion of green energy within the scientific community. AC from olive mill solid waste was prepared by chemical impregnation with zinc chloride. The synthesized materials were characterized by different techniques (XRD, BET, SEM-EDX, FT-IR, UV-Vis) to study their physicochemical properties. The FeCr/AC composite was used to study the photo-Fenton degradation of different types of model pollutants. The photo-Fenton performance of MB was evaluated by modifying several factors, such as contact time, catalyst loading, dye concentration, and pH, on the degradation of MB. Additionally, the innovative design of our composite promotes higher reusability and stability, addressing the key challenges in photocatalytic applications. The degradation performance was further tested using methyl orange (MO) and the tetracycline (TCH) antibiotic to validate the composite’s versatility across different pollutant classes.

2. Materials and Methods

2.1. Chemicals

All reagents utilized in this study were of analytical grade and employed without any additional purification. Zinc chloride (ZnCl2), iron nitrate nonahydrate (Fe(NO3)3·9H2O), chromium nitrate nonahydrate (Cr(NO3)3·9H2O), ammonia (NH3, 25% w/w), hydrochloric acid (37% HCl), sodium hydroxide (NaOH), hydrogen peroxide (H2O2, 35% w/w, density 1.13 g/mL), methylene blue (C16H18N3SCl, MW: 373.88 g/mole), methyl orange (C14H14N3NaO3S, MW: 327.33 g/mole), and tetracycline hydrochloride (C22H24N2O8·HCl, MW: 480.90) were all procured from Sigma-Aldrich (Beirut, Lebanon).

2.2. Preparation of Activated Carbon Support

The collected olive-processing residues were initially washed using deionized water and then crushed and sieved to obtain a uniform particle size. The resulting particles were impregnated with a ZnCl2 solution (impregnation ratio of ZnCl2 to OMSW = 2:1) for 24 h at an ambient temperature. Following impregnation, the sample was oven-dried at 110 °C for 24 h. The dried mixture was transferred to a sealed crucible and subjected to thermal treatment in a programmable muffle furnace at 500 °C for 2 h with a heating rate of 5 °C min−1. The obtained AC was then cooled and washed with 0.1 M HCl and distilled water to eliminate any residual chlorides and mineral matter. Finally, the AC was dried again at 110 °C and stored for further use.

2.3. Preparation of Composite Mixture (FeCr/AC)

Chromium ferrite supported on activated carbon (AC), derived from OMSW, was prepared using the hydrothermal method. First, a known mass for each metal nitrate solution, Fe(NO3)3·9H2O and Cr(NO3)3·9H2O, was dissolved in 10 mL of distilled water in two separate beakers to achieve final percentages of 8% iron and 4% chromium in the final prepared catalyst. The two solutions were then mixed together, after which one gram of AC was added to them. Following this, the suspension was sonicated for 1 h to obtain a uniform suspension and to ensure the proper dispersion of iron and chromium precursors on the AC’s surface. Then, an ammonia solution (25% w/w) was added to the suspension until the pH reached 10. The resulting solid was collected through filtration, thoroughly rinsed with deionized water, and then left to dry overnight at 110 °C. Subsequently, the dried powders were subjected to calcination at 500 °C for 2 h using a heating rate of 5 °C per minute.

2.4. Characterization Techniques

Various analytical tools were employed to characterize the synthesized materials. The crystal structure and phase composition of both AC and FeCr/AC composites were examined via X-ray diffraction (XRD) using a Bruker D8 Advance diffractometer (Karlsruhe, Germany) equipped with a Cu Kα radiation source (λ = 1.5418 Å), operated at 40 kV and 40 mA, and scanned over a 2θ range of 10–60°. The textural properties, including surface area and porosity, were investigated by nitrogen adsorption–desorption measurements at 77 K using a Micromeritics Gemini VII 2390p analyzer (Norcross, GA, USA). Before analysis, the samples were degassed under vacuum at 120 °C overnight. The specific surface area (SBET) was derived using the BET method over a relative pressure range of 0.01–0.1. Pore distribution was determined by the Barrett–Joyner–Halenda (BJH) method applied to the desorption curve, while micropore volume was obtained using the t-plot technique.
To assess the surface morphology and elemental content, scanning electron microscopy (SEM) analysis was conducted using a Tescan MIRA3 microscope fitted with an EDX detector (Hillsboro, OR, USA). A small amount of catalyst powder was affixed to a carbon-coated aluminum stub and coated with approximately 5 nm of gold by sputtering. Functional groups present on the AC and FeCr/AC surfaces were identified using FTIR spectroscopy (Thermo Nicolet Summit-LITE, Darmstadt, Germany), scanning in the 400–4000 cm−1 range at a resolution of 4 cm−1. The optical properties and band gap of FeCr/AC were assessed through diffuse reflectance spectroscopy (DRS) with a Perkin Elmer UV-Vis-NIR spectrophotometer (Lambda 1050+ model, Thermo Fisher Scientific, Karlsruhe, Germany). To study the leaching of metals, an atomic absorption spectroscopy machine (AAS) was employed using a flame-based AAS instrument (Thermo Fisher Scientific ICE 3000 Series spectrometer, Karlsruhe, Germany). Calibration was performed using standard solutions for each metal. The detection limits were determined to be <12 ppm for Fe and <8 ppm for Cr. Each measurement was performed in triplicate, and the results are reported as the average of the three runs to ensure accuracy and reproducibility.

2.5. Process for the Photo-Fenton Degradation Experiment

The photocatalytic degradation experiment was conducted in a 250 mL beaker under visible light using a 25 W white LED lamp. Different pollutants, including cationic dye (methylene blue, MB), anionic dye (methyl orange, MO), and pharmaceutical (tetracycline hydrochloride), were taken as model pollutants. For the degradation of the MB pollutant, varying doses of FeCr/AC were suspended in 100 mL of MB solution with an initial concentration of 20 ppm. The suspension was stirred in the dark for 30 min before irradiation to achieve an equilibrium state between the pollutant and catalyst. Next, the lamp was turned on and 0.25 mL H2O2 was introduced into the beaker. Samples were then withdrawn at different time intervals (t = 0 to t = 180 min). These samples were filtered using a 0.45 µm syringe filter. The concentration of residual pollutants in the filtered solution was recorded using a UV-visible spectrophotometer by measuring the absorbance at maximum wavelengths of 664, 464, and 365 nm for MB, MO, and TCH, respectively. The degradation efficiency was computed using Equation (1).
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: initial pollutant concentration (mg L−1) and Ct: pollutant concentration at time t (mg L−1).

3. Results

3.1. X-Ray Diffraction

Figure 1 displays the wide-angle XRD patterns of the AC support and the FeCr/AC composite. For AC, a broad peak appears within the 2θ range of 20–30°, which corresponds to the (002) reflection plane typically associated with hexagonal graphitic carbon structures [53]. FeCr/AC showed sharp and intense peaks at 2θ angles of 30°, 35.5°, 43°, 53.5°, 57°, 62.5°, and 74°, which can be indexed to planes (220), (311), (400), (422), (511), (440), and (533), respectively [54]. These peaks are in good agreement with the cubic spinel structure of metal ferrites with space group Fd-3m according to JCPDS card no.10-0325 [55]. No other impurities were detected in the XRD pattern, which indicates the high purity of the synthesized material. Therefore, CrFe2O4 nanoparticles are effectively incorporated into AC.
The crystallite size of FeCr/AC is determined from the Scherrer equation given below:
D =       K λ / β c o s θ    
where D represents the crystallite size, K is the shape factor (taken as 0.9 for cubic structure), λ is the X-ray wavelength (Cu Kα, 0.15418 nm), β is the full width at half maximum (FWHM) of the most intense diffraction peak (in radians), and   θ is the Bragg angle.
The FeCr/AC crystallite size was determined to be 12 nm at the 311 diffraction plane. This value is consistent with the previously reported crystallite sizes of CoFe2O4 [56].

3.2. Textural Properties

The surface area and pore structure of the pure AC support and FeCr/AC were assessed by nitrogen adsorption–desorption isotherms at 77 K. The isotherms are displayed in Figure 2. The adsorption isotherm of AC exhibited a type I isotherm, typical for microporous materials. This is accompanied by a hysteresis loop of type H4 in the relative pressure range of 0.4–0.8 P/P0. The prominent hysteresis loop, which is a property of mesoporous materials, is caused by capillary condensation in slit-shaped pores. Conversely, the sorption isotherm of modified FeCr/AC revealed a type IV isotherm associated with the H1 hysteresis loop, typical for mesoporous materials [57].
The textural parameters, including BET surface area (SBET), total pore volume (Vp), pore diameter (Dp), and micropore volume (Vµp), of the prepared AC and FeCr/AC are listed in Table 1. The AC exhibited a high SBET 1148 m2 g−1 and a Vp of 0.6 cm3 g−1. After the deposition of Cr and Fe, there was a significant reduction in the SBET and Vp to 222 m2 g−1 and 0.3 cm3 g−1, respectively. The reduction in surface area after metal deposition suggests that the nanoparticles are well incorporated within the AC pores.

3.3. Fourier Transform Infrared (FT-IR)

Figure 3 presents the FTIR spectra of AC and FeCr/AC. For both spectra, there exists a wide and large band at 3365 cm−1, which refers to the stretching vibration band of OH in carbonyl, hydroxyl, or adsorbed H2O molecules. The intensity of this band decreases following the incorporation of metals in the FeCr/AC spectrum. The stretching vibration band of C=C in the benzene ring is confirmed by the peak observed at 1570 cm−1 in the spectrum of AC [58], while, in FeCr/AC, a band is detected at 1620 cm−1, which can be assigned to the asymmetric stretching vibration of C=O [59] or the bending vibration of H2O [60]. The absorption band at 946 cm−1 in FeCr/AC confirms the presence of an aromatic group [61]. Additionally, the band observed at 2970 cm−1 corresponds to the symmetric stretching vibration of C-H [62]. A sharp and intense absorption band appears at 570 cm−1 in the spectrum of FeCr/AC, which was absent in the AC spectrum. This newly formed band can be attributed to the stretching vibration of M-O at the tetrahedral sites [63].

3.4. Scanning Electron Microscope (SEM)

The surface morphology and elemental distribution of the synthesized samples were analyzed through SEM imaging combined with EDX spectroscopy, as presented in Figure 4. The SEM micrograph of AC displays a highly porous and rough surface morphology, characterized by an irregular and heterogeneous structure with numerous cavities and voids. These porous characteristics contribute to a high surface area, which enhances the adsorption capacity and provides suitable support for metal dispersion in photocatalytic applications. In contrast, Figure 4b shows a more compact and dense surface with reduced porosity. This is likely due to pore filling and surface coverage by the dispersed Fe and Cr metals. Additionally, the presence of large, flake-like structures coated with a rough, crust-like layer indicates that successful deposition of chromium ferrite nanoparticles across the AC’s surface.
For a further analysis of elemental composition, the relevant regions of AC and FeCr/AC were detected by EDS mapping. As shown in Figure 4c, AC comprises mainly two components (carbon and oxygen) with relative abundances of 89.16 and 10.84 wt.%, respectively. However, after modification with chromium ferrite, new peaks for Fe and Cr are observed in Figure 4d with relative abundances of 7.79 and 3.9 wt.%, respectively. These findings confirm the successful incorporation of Fe and Cr into the AC material.

3.5. UV-Vis Spectroscopy

In order to evaluate the optical characteristics and estimate the band gap of the synthesized materials, DRS spectra were recorded, and the findings are illustrated in Figure 5a. Spinel ferrite nanoparticles exhibit light absorption within the visible range due to electrons’ transition from the O-2p level to the Fe-3d level [64]. The optical behavior evident in the DRS data for ferrites is attributed to electron excitation occurring from the valence band (VB) to the conduction band (CB). This excitation energy is directly linked to the ferrite band gap. In general, the Kubelka–Munk method is utilized with DRS data to compute the band gap (Equation (3)).
F R = α = ( 1 R ) 2 2 R            
where F(R) denotes the Kubelka–Munk function, α refers to the absorption coefficient, and R represents the reflectance.
The Tauc plot equation is represented as follows in Equation (4):
( F R h v ) 1 n = A ( h v E B G )                                
In this equation, h is Planck’s constant, A is a proportionality factor, v is the frequency of incident photons, and EBG indicates the band gap energy. The exponent n varies based on the type of electronic transition. Depending on the type of transition, n can be 2 for indirect or ½ for direct transitions. As reported in the literature, spinel ferrites have a direct band gap [65]. Therefore, in our case, n is equal to ½. The Tauc plot of FeCr/AC is shown in Figure 5b. As is revealed, the band gap energy for FeCr/AC is 1.9 eV. Table 2 presents a comparison of the band gap energies of various metal ferrite photocatalysts reported in the literature. Among the catalysts examined, our prepared CrFe2O4/AC photocatalyst showed the lowest band gap (1.90 eV), indicating better potential for visible-light-driven photocatalytic applications.

3.6. Photo-Fenton Catalytic Activity of FeCr/AC

Several experimental conditions were tested to compare the decolorization efficiency of MB, and the results are presented in Figure 6. These conditions included visible light only, FeCr/AC with visible light, FeCr/AC with H2O2 and visible light, and FeCr/AC in the dark. The blank experiment, conducted under visible-light irradiation and in the absence of FeCr/AC catalyst and H2O2, revealed only 1.6% MB degradation after 120 min, indicating that the photolysis reaction was negligible. Prior to illumination, a dark experiment was performed for 180 min to study the adsorption of MB onto FeCr/AC. Equilibrium between adsorption and desorption was achieved within 30 min, resulting in a 25% removal of MB attributed to the adsorption process. Interestingly, the results reveal a complete decolorization of MB (97.56%) after 120 min when FeCr/AC is irradiated with a 25W LED lamp in the presence of H2O2. This result is significantly better than that obtained using FeCr/AC alone (35%) in the presence of visible light without adding H2O2. Thus, H2O2 plays a pivotal function in the degradation mechanism of MB and facilitates oxidation reactions in photocatalytic activities [71]. The aforementioned findings show that the combination of FeCr/AC, H2O2 and visible light is necessary for the effective photo-Fenton degradation of MB dye.

3.7. Impact of Catalyst Amount on MB Degradation

The influence of FeCr/AC dosage on MB degradation was investigated using varying FeCr/AC quantities of 125, 250, and 400 ppm at a fixed MB concentration of 20 ppm, 0.25 mL H2O2 (35%), and a near-neutral pH of 6.47 (Figure 7). The results reveal an improvement in MB degradation from 54.2% to 86.8% as the catalyst amount increased from 125 ppm to 250 ppm at 60 min under LED irradiation. This enhancement can be attributed to the increased availability of active sites on the FeCr/AC surface, which improves its light-harvesting capability. As more photons are absorbed, more charge carriers are generated, resulting in an increase in the number of hydroxyl radicals formed and, consequently, an improvement in the degradation of MB. However, there was almost no improvement in the degradation of MB as the catalyst amount exceeded 250 ppm, possibly due to the agglomeration of nanoparticles at high catalyst doses. Thus, in our study, the optimum dosage of FeCr is determined to be 250 ppm.

3.8. Impact of Initial MB Concentration

The effect of the starting dye concentration on MB degradation was investigated by varying MB levels between 10 and 100 ppm, with a fixed catalyst amount of 250 ppm, 0.25 mL H2O2, and a near-neutral pH of 6.47. The degradation of MB reduced from 92% to 43.6% as the MB concentration rose from 10 ppm to 100 ppm (Figure 8). At elevated dye concentrations, multiple factors contributed to the observed decrease in degradation efficiency. First, the increased optical density of the solution reduces light penetration, limiting the activation of the photocatalyst surface [72]. Second, a higher substrate concentration leads to competition for available hydroxyl radicals, since the radical generation rate remains constant while more dye molecules are present. Third, the accumulation of intermediate degradation products may also consume hydroxyl radicals [73], further decreasing the effective radical availability for parent dye molecules. These combined effects result in a lower degradation rate constant at higher methyl orange concentrations.

3.9. Impact of Initial pH

The influence of solution pH on the degradation of MB by FeCr/AC was examined using three distinct pH values (3, 6.47, and 10) with 250 ppm FeCr/AC, 20 ppm MB, and 0.25 mL H2O2. The results are represented in Figure 9. The degradation of MB was the highest at neutral pH and reached 87% in 60 min, indicating that neutral conditions are favorable for the photo-Fenton degradation of MB. However, a decline in MB degradation was observed under acidic and basic conditions, at 29% and 50%, respectively. Although homogeneous Fenton systems typically exhibit an optimal performance at pH ~3 due to the efficient generation of hydroxyl radicals via Fe2+/Fe3+ cycling [74], our system is based on a heterogeneous ferrite catalyst. In such systems, surface interactions play a key role in degradation efficiency. The reduction in MB degradation at an acidic pH was expected due to the electrostatic repulsion between MB cationic dye and the surface of FeCr/AC, which also becomes more positively charged under highly acidic conditions, thereby hindering effective adsorption and interaction. In contrast, in alkaline conditions, despite the negatively charged environment, the degradation of MB was reduced to 50%. This reduction is attributed to the production of weak oxidant ferryl ions (FeO2+), which are formed at pH > 5 according to Equation (5) [75]. Additionally, in a basic medium, a hydrogen peroxide oxidant is unstable and its degradation produces water and oxygen (Equation (6)). In this study, the optimum pH was determined to be 6.47.
F e 2 + + H 2 O 2   F e O 2 + + H 2 O
2 H 2 O 2 2 H 2 O + O 2            

3.10. Impact of Oxidant Dosage on MB Degradation

In the photo-Fenton oxidation process, H2O2 plays a key role in generating hydroxyl radicals, which participate in the mineralization of MB dyes. The impact of H2O2 amount was evaluated using different dosages of H2O2 ranging from 0.25 mL to 1.0 mL at 250 ppm FeCr/AC, a MB concentration of 20 ppm, and a neutral pH. As shown in Figure 10, an almost complete degradation of MB (97%) was achieved when the H2O2 concentration was 0.25 mL. However, the degradation decreased gradually from 97% to 86.6% as the dosage of H2O2 increased to 1 mL. This reduction is due to the formation of hydroperoxyl radicals (HO2.) at excess concentrations, which act as scavengers of hydroxyl radicals (Equation (7)) [76]. Therefore, in our study, the addition of only 0.25 mL of H2O2 was sufficient to optimize the degradation of MB.
H 2 O 2     + · O H     H 2 O     + · O O H          

3.11. Kinetic Study of MB

The degradation rate of MB can be evaluated using kinetic parameters. To ascertain the kinetics of the degradation of MB, the experiment was performed using 250 ppm FeCr/AC, a pH of 6.47, and a time interval up to 180 min under visible LED light. The degradation kinetics of MB at different concentrations (10, 20, and 100 ppm) were studied by plotting ln (C0/Ct) versus time (min). The rate constant (kapp) was determined from the slope of the linear graph (Equation (7)). Table 3 presents the values of the decomposition rate constant (k) and the half-life time (t1/2) for the different initial concentrations of MB (Equation (8)). As shown in Table 3, the Langmuir–Hinshelwood pseudo-first-order kinetic model fits the experimental data well, with R2 > 0.97. The calculated rate constants for different concentrations of MB were 0.0297 min−1, 0.0258 min−1, and 0.0077 min−1 for 10 ppm, 20 ppm, and 100 ppm concentrations of MB, respectively. The reaction rate constant reduced from 0.0297 min−1 to 0.0077 min−1 with increasing MB concentration from 10 ppm to 100 ppm. As previously discussed, the observed decrease in the degradation rate at higher dye concentrations is attributed to the combined effects of reduced light penetration, competition for hydroxyl radicals, and radical scavenging by intermediate products (see Section 3.8). Conversely, the half-life of MB increased from 23.33 min−1 to 89.78 min−1 as the concentration of MB increased from 10 ppm to 100 ppm.
ln C 0 C t   = k a p p t              
t 1 / 2 = l n 2 k a p p
where C0: MB initial concentration (mg L−1), Ct: final concentration (mg L−1), kapp: first-order rate constant (min−1), t: reaction time (min), and t1/2: half-life time (min).

3.12. FeCr/AC Recyclability and Leaching Test

The reusability of catalysts is a crucial requirement for industrial applications. To assess the reusability of FeCr/AC under visible light, the catalyst was tested over three consecutive cycles. Following each cycle, the catalyst was recovered from the MB solution, washed with deionized water, and then dried at 100 °C prior to its reuse in subsequent reactions. The photocatalytic efficiency of FeCr/AC remained high, starting at 98% in cycle 1, achieving 92.76% in cycle 2, and maintaining a robust 85.41% in cycle 3. These results validate the reusability and durability of the prepared photocatalyst nanocomposite. Moreover, the leaching of Fe into the solution as detected by atomic absorption spectroscopy was 0.1 ppm after the third cycle, while Cr leaching remained below the detection limit of the instrument.

3.13. Possible Mechanism for MB Degradation by FeCr/AC

When chromium ferrite nanoparticles supported on AC absorb visible light, electron-hole pairs are generated. The electrons (e) in the valence band (VB) of CrFe2O4/AC are promoted to the conduction band (CB), leaving behind holes (h+) in the VB (Equation (10)). The photogenerated electrons in the conduction band can reduce dissolved O2 molecules, forming superoxide anions (O2•), which can contribute to additional degradation pathways alongside classical Fenton reactions (Equation (11)), as similarly reported for semiconductor-based hybrid systems. These anions can directly degrade MB molecules or react with H2O2 to produce hydroxyl radicals (OH•) (Equation (12)). On the contrary, the h+ in the VB may react with H2O molecules to form OH• (Equation (13)). These radicals are potent oxidizers responsible for MB degradation (Equation (14)).
C r F e 2 O 4 / A C + h v e C B + h V B +
e C B + O 2 O 2 .
H 2 O 2 + O 2 . O H . + O H + O 2
h V B + + H 2 O O H . + H +
M B   d y e + O H .   o r   O 2 . C O 2 + H 2 O + m i n e r a l i z a t i o n   p r o d u c t s

3.14. Degradation of MO and TCH

The photocatalytic performance of FeCr/AC under visible LED irradiation is investigated for the degradation of MO and TC using 250 ppm FeCr/AC, 20 ppm MO, 20 ppm TCH, and 0.5 mL H2O2, without modifying the solution pH (5.47 for MO and 4.4 for TCH). Figure 11 presents the degradation results for both pollutants. Dark experiments were also performed for 180 min to assess the adsorption of MO and TCH. The results reveal no adsorption of MO and around 17% adsorption of TCH. According to Figure 11, maximum removal rates of 88% for MO and 97% for TC can be observed.
The Langmuir–Hinshelwood first-order kinetic model is suitable to fit the degradation of MO and TCH. The first-order kinetic parameters k, R2, and t1/2 for MO and TCH are mentioned in Table 4. The obtained k1 values for TCH and MO are 0.0225 min−1 and 0.0115 min−1, respectively.

3.15. Comparison with the Literature

The catalytic performance of the FeCr/AC composite is evaluated for the degradation of MB and compared with other ferrites-based photocatalysts reported in the literature. As revealed in Table 5, the FeCr/AC photocatalyst developed in this study exhibits notable advantages over previously reported ferrite-based catalysts for the degradation of MB. The FeCr/AC composite achieved a high removal efficiency of 97.56% using lower catalyst loading (0.25 g L−1) under a near-neutral pH (6.47) within 120 min of irradiation. The use of a low-power LED light source further underscores the energy efficiency and cost-effectiveness of the process under ambient conditions. Moreover, the utilization of AC derived from OMSW as a catalyst support not only reduces material costs but also enhances the environmental sustainability and structural stability of the photocatalyst, which are essential for practical wastewater treatment applications.

4. Conclusions

A FeCr/AC composite was prepared by the hydrothermal synthesis process and used to study the degradation of MB dye and different pollutants (MO and TCH) using a heterogeneous photo-Fenton process under visible 25 W LED illumination. The physicochemical characteristics of FeCr/AC were investigated using techniques including XRD, BET, SEM, EDX, FT-IR, and DRS. The characterization results confirm the successful incorporation of chromium ferrite nanoparticles into AC support. Moreover, the band gap of FeCr/AC, determined from DRS spectra, was 1.9 eV, demonstrating its efficiency in visible-light photocatalytic applications. The prepared FeCr/AC catalyst exhibited high efficacy in the decolorization of MB, achieving 97% removal efficiency within 120 min in a neutral environment using only 0.25 mL H2O2 under visible LED illumination. The degradation kinetics of MB using FeCr/AC followed a pseudo-first-order model, yielding a rate constant of 0.0258 min−1 at an initial concentration of 20 ppm. Furthermore, FeCr/AC demonstrated notable reusability and stability, with only a minor reduction in degradation the performance observed after three consecutive cycles. Additionally, FeCr/AC showed high removal rates for MO and TCH, 88% and 97%, respectively, within 180 min under the following specified reaction conditions: 250 ppm FeCr, 0.5 mL H2O2, and an unmodified pH. The corresponding rate constants of MO and TCH were 0.0115 and 0.0225 min−1, respectively. Therefore, the prepared FeCr/AC catalyst shows promise for large-scale application in the effective treatment of contaminated water containing diverse pollutants.

Author Contributions

Conceptualization, all authors; methodology, all authors; software, M.H. (Malak Hamieh); validation, all authors; formal analysis, all authors; investigation, M.H. (Mohammad Hammoud), S.A.K. and K.C.; resources, J.T. and M.H. (Malak Hamieh); data curation, M.H. (Mohammad Hammoud) and J.T.; writing—original draft preparation, all authors; writing—review and editing, M.H. (Malak Hamieh), J.T. and T.H.; visualization, all authors; supervision, J.T.; project administration, J.T.; funding acquisition, J.T. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

All data generated or analyzed during this study are included in this published article.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. World Economic Forum. 5 Lessons for the Future of Water. 2020. Available online: https://www.weforum.org/agenda/2020/04/covid-19-water-what-can-we-learn/ (accessed on 17 September 2024).
  2. Zulfiqar, M.; Samsudin, M.F.R.; Sufian, S. Modelling and optimization of photocatalytic degradation of phenol via TiO2 nanoparticles: An insight into response surface methodology and artificial neural network. J. Photochem. Photobiol. A Chem. 2019, 384, 112039. [Google Scholar] [CrossRef]
  3. Pavithra, K.G.; Kumar, P.S.; Jaikumar, V.; Rajan, P.S. Removal of colorants from wastewater: A review on sources and treatment strategies. J. Ind. Eng. Chem. 2019, 75, 1–19. [Google Scholar] [CrossRef]
  4. Vanhulle, S.; Trovaslet, M.; Enaud, E.; Lucas, M.; Taghavi, S.; Van der Lelie, D.; Van Aken, B.; Foret, M.; Onderwater, R.C.A.; Wesenberg, D.; et al. Genotoxicity Reduction During a Combined Ozonation/Fungal Treatment of Dye-Contaminated Wastewater. Environ. Sci. Technol. 2008, 42, 584–589. [Google Scholar] [CrossRef] [PubMed]
  5. Das, T.R.; Patra, S.; Madhuri, R.; Sharma, P.K. Bismuth oxide decorated graphene oxide nanocomposites synthesized via sonochemical assisted hydrothermal method for adsorption of cationic organic dyes. J. Colloid Interface Sci. 2018, 509, 82–93. [Google Scholar] [CrossRef] [PubMed]
  6. Hafiz, M.; Hassanein, A.; Talhami, M.; Al-Ejji, M.; Hassan, M.K.; Hawari, A.H. Magnetic nanoparticles draw solution for forward osmosis: Current status and future challenges in wastewater treatment. J. Environ. Chem. Eng. 2022, 10, 108955. [Google Scholar] [CrossRef]
  7. Aslan, S.; Şirazi, M. Adsorption of Sulfonamide Antibiotic onto Activated Carbon Prepared from an Agro-industrial By-Product as Low-Cost Adsorbent: Equilibrium, Thermodynamic, and Kinetic Studies. Water Air Soil Pollut. 2020, 231, 222. [Google Scholar] [CrossRef]
  8. Omri, A.; Wali, A.; Benzina, M. Adsorption of bentazon on activated carbon prepared from Lawsonia inermis wood: Equilibrium, kinetic and thermodynamic studies. Arab. J. Chem. 2016, 9, S1729–S1739. [Google Scholar] [CrossRef]
  9. Rosal, R.; Rodríguez, A.; Perdigón-Melón, J.A.; Petre, A.; García-Calvo, E.; Gómez, M.J.; Agüera, A.; Fernández-Alba, A.R. Occurrence of emerging pollutants in urban wastewater and their removal through biological treatment followed by ozonation. Water Res. 2010, 44, 578–588. [Google Scholar] [CrossRef]
  10. Ali, T.; Tripathi, P.; Azam, A.; Raza, W.; Ahmed, A.S.; Ahmed, A.; Muneer, M. Photocatalytic performance of Fe-doped TiO2 nanoparticles under visible-light irradiation. Mater. Res. Express 2017, 4, 15022. [Google Scholar] [CrossRef]
  11. Karthik, C.; Swathi, N.; Pandi Prabha, S.; Caroline, D.G. Green synthesized rGO-AgNP hybrid nanocomposite—An effective antibacterial adsorbent for photocatalytic removal of DB-14 dye from aqueous solution. J. Environ. Chem. Eng. 2020, 8, 103577. [Google Scholar] [CrossRef]
  12. McMullan, G.; Meehan, C.; Conneely, A.; Kirby, N.; Robinson, T.; Nigam, P.; Banat, I.M.; Marchant, R.; Smyth, W.F. Microbial decolourisation and degradation of textile dyes. Appl. Microbiol. Biotechnol. 2001, 56, 81–87. [Google Scholar] [CrossRef]
  13. Emamjomeh, M.M.; Sivakumar, M. Review of pollutants removed by electrocoagulation and electrocoagulation/flotation processes. J. Environ. Manag. 2009, 90, 1663–1679. [Google Scholar] [CrossRef]
  14. Martínez-Huitle, C.A.; Rodrigo, M.A.; Sirés, I.; Scialdone, O. Single and Coupled Electrochemical Processes and Reactors for the Abatement of Organic Water Pollutants: A Critical Review. Chem. Rev. 2015, 115, 13362–13407. [Google Scholar] [CrossRef] [PubMed]
  15. Momina; Ahmad, K. Feasibility of the adsorption as a process for its large scale adoption across industries for the treatment of wastewater: Research gaps and economic assessment. J. Clean. Prod. 2023, 388, 136014. [Google Scholar] [CrossRef]
  16. Kim, S.-H.; Kim, D.-S.; Moradi, H.; Chang, Y.-Y.; Yang, J.-K. Highly porous biobased graphene-like carbon adsorbent for dye removal: Preparation, adsorption mechanisms and optimization. J. Environ. Chem. Eng. 2023, 11, 109278. [Google Scholar] [CrossRef]
  17. Dhamorikar, R.S.; Lade, V.G.; Kewalramani, P.V.; Bindwal, A.B. Review on integrated advanced oxidation processes for water and wastewater treatment. J. Ind. Eng. Chem. 2024, 138, 104–122. [Google Scholar] [CrossRef]
  18. Sirés, I.; Brillas, E.; Oturan, M.A.; Rodrigo, M.A.; Panizza, M. Electrochemical advanced oxidation processes: Today and tomorrow. A review. Environ. Sci. Pollut. Res. 2014, 21, 8336–8367. [Google Scholar] [CrossRef]
  19. Ahmed, S.; Rasul, M.G.; Brown, R.; Hashib, M.A. Influence of parameters on the heterogeneous photocatalytic degradation of pesticides and phenolic contaminants in wastewater: A short review. J. Environ. Manag. 2011, 92, 311–330. [Google Scholar] [CrossRef]
  20. Sun, B.; Li, H.; Li, X.; Liu, X.; Zhang, C.; Xu, H.; Zhao, X.S. Degradation of Organic Dyes over Fenton-Like Cu2O-Cu/C Catalysts. Ind. Eng. Chem. Res. 2018, 57, 14011–14021. [Google Scholar] [CrossRef]
  21. Susanti, Y.D.; Saleh, R. Efficient photo-, sono-, and sonophoto-fenton-like degradation of the organic pollutant methylene blue using a BiFeO3/graphene composite. IOP Conf. Ser. Mater. Sci. Eng. 2020, 763, 012064. [Google Scholar] [CrossRef]
  22. Vaishnave, P.; Ameta, G.; Kumara, A.; Sharma, S.; Ameta, S.C. Sono-photo-Fenton and photo-Fenton degradation of methylene blue: A comparative study. J. Indian Chem. Soc. 2011, 88, 397–403. [Google Scholar] [CrossRef]
  23. Halfadji, A.; Naous, M.; Kharroubi, K.N.; Belmehdi, F.E.Z.; Aoudia, H. Facile prepared Fe3O4 nanoparticles as a nano-catalyst on photo-fenton process to remediation of methylene blue dye from water: Characterization and optimization. Glob. NEST J. 2024, 26, 1–6. [Google Scholar] [CrossRef]
  24. Zhu, Y.; Zhu, R.; Yan, L.; Fu, H.; Xi, Y.; Zhou, H.; Zhu, G.; Zhu, J.; He, H. Visible-light Ag/AgBr/ferrihydrite catalyst with enhanced heterogeneous photo-Fenton reactivity via electron transfer from Ag/AgBr to ferrihydrite. Appl. Catal. B Environ. 2018, 239, 280–289. [Google Scholar] [CrossRef]
  25. Jiang, J.; Gao, J.; Li, T.; Chen, Y.; Wu, Q.; Xie, T.; Lin, Y.; Dong, S. Visible-light-driven photo-Fenton reaction with α-Fe2O3/BiOI at near neutral pH: Boosted photogenerated charge separation, optimum operating parameters and mechanism insight. J. Colloid Interface Sci. 2019, 554, 531–543. [Google Scholar] [CrossRef]
  26. Ahmed, Y.; Yaakob, Z.; Akhtar, P. Degradation and mineralization of methylene blue using a heterogeneous photo-Fenton catalyst under visible and solar light irradiation. Catal. Sci. Technol. 2016, 6, 1222–1232. [Google Scholar] [CrossRef]
  27. Wang, L.; Zhou, C.; Yuan, Y.; Jin, Y.; Liu, Y.; Jiang, Z.; Li, X.; Dai, J.; Zhang, Y.; Siyal, A.A.; et al. Catalytic degradation of crystal violet and methyl orange in heterogeneous Fenton-like processes. Chemosphere 2023, 344, 140406. [Google Scholar] [CrossRef] [PubMed]
  28. Tryba, B.; Piszcz, M.; Grzmil, B.; Pattek-Janczyk, A.; Morawski, A.W. Photodecomposition of dyes on Fe-C-TiO2 photocatalysts under UV radiation supported by photo-Fenton process. J. Hazard. Mater. 2009, 162, 111–119. [Google Scholar] [CrossRef]
  29. Liu, S.-Q.; Feng, L.-R.; Xu, N.; Chen, Z.-G.; Wang, X.-M. Magnetic nickel ferrite as a heterogeneous photo-Fenton catalyst for the degradation of rhodamine B in the presence of oxalic acid. Chem. Eng. J. 2012, 203, 432–439. [Google Scholar] [CrossRef]
  30. Diao, Y.; Yan, Z.; Guo, M.; Wang, X. Magnetic multi-metal co-doped magnesium ferrite nanoparticles: An efficient visible light-assisted heterogeneous Fenton-like catalyst synthesized from saprolite laterite ore. J. Hazard. Mater. 2018, 344, 829–838. [Google Scholar]
  31. Anchieta, C.G.; Severo, E.C.; Rigo, C.; Mazutti, M.A.; Kuhn, R.C.; Muller, E.I.; Flores, E.M.M.; Moreira, R.F.P.M.; Foletto, E.L. Rapid and facile preparation of zinc ferrite (ZnFe2O4) oxide by microwave-solvothermal technique and its catalytic activity in heterogeneous photo-Fenton reaction. Mater. Chem. Phys. 2015, 160, 141–147. [Google Scholar] [CrossRef]
  32. Firouzeh, N.; Paseban, A.; Ghorbanian, M.; Asadzadeh, S.N.; Amani, A. Recyclable Nano-Magnetic CoFe2O4: A Photo-Fenton Catalyst for Efficient Degradation of Reactive Blue 19. Bionanoscience 2024, 14, 4481–4492. [Google Scholar] [CrossRef]
  33. Fahad Almojil, S.; Ning, J.; Ibrahim Almohana, A. Photo-Fenton process for degradation of methylene blue using copper ferrite@sepiolite clay. Inorg. Chem. Commun. 2024, 166, 12623. [Google Scholar] [CrossRef]
  34. Qin, L.; Wang, Z.; Fu, Y.; Lai, C.; Liu, X.; Li, B.; Liu, S.; Yi, H.; Li, L.; Zhang, M.; et al. Gold nanoparticles-modified MnFe2O4 with synergistic catalysis for photo-Fenton degradation of tetracycline under neutral pH. J. Hazard. Mater. 2021, 414, 125448. [Google Scholar] [CrossRef] [PubMed]
  35. Jadhav, S.A.; Somvanshi, S.B.; Khedkar, M.V.; Patade, S.R.; Jadhav, K.M. Magneto-structural and photocatalytic behavior of mixed Ni–Zn nano-spinel ferrites: Visible light-enabled active photodegradation of rhodamine B. J. Mater. Sci. Mater. Electron. 2020, 31, 11352–11365. [Google Scholar] [CrossRef]
  36. Mendonça, M.H.; Godinho, M.I.; Catarino, M.A.; da Silva Pereira, M.I.; Costa, F.M. Preparation and characterization of spinel oxide ferrites suitable for oxygen evolution anodes. Solid State Sci. 2002, 4, 175–182. [Google Scholar] [CrossRef]
  37. Valente, F.; Astolfi, L.; Simoni, E.; Danti, S.; Franceschini, V.; Chicca, M.; Martini, A. Nanoparticle drug delivery systems for inner ear therapy: An overview. J. Drug Deliv. Sci. Technol. 2017, 39, 28–35. [Google Scholar] [CrossRef]
  38. Casbeer, E.; Sharma, V.K.; Li, X. Z Synthesis and photocatalytic activity of ferrites under visible light: A review. Sep. Purif. Technol. 2012, 87, 1–14. [Google Scholar] [CrossRef]
  39. Zhou, L.; Ji, L.; Ma, P.-C.; Shao, Y.; Zhang, H.; Gao, W.; Li, Y. Development of carbon nanotubes/CoFe2O4 magnetic hybrid material for removal of tetrabromobisphenol A and Pb(II). J. Hazard. Mater. 2014, 265, 104–114. [Google Scholar] [CrossRef]
  40. Rooygar, A.A.; Mallah, M.H.; Abolghasemi, H.; Safdari, J. New ‘magmolecular’ process for the separation of antimony(III) from aqueous solution. J. Chem. Eng. Data 2014, 59, 3545–3554. [Google Scholar] [CrossRef]
  41. Goswami, M.; Phukan, P. Enhanced adsorption of cationic dyes using sulfonic acid modified activated carbon. J. Environ. Chem. Eng. 2017, 5, 3508–3517. [Google Scholar] [CrossRef]
  42. Zabihi, M.; Khorasheh, F.; Shayegan, J. Supported copper and cobalt oxides on activated carbon for simultaneous oxidation of toluene and cyclohexane in air. RSC Adv. 2015, 5, 5107–5122. [Google Scholar] [CrossRef]
  43. Bouriche, R.; Tazibet, S.; Boutillara, Y.; Melouki, R.; Benaliouche, F.; Boucheffa, Y. Characterization of Titanium (IV) Oxide Nanoparticles Loaded onto Activated Carbon for the Adsorption of Nitrogen Oxides Produced from the Degradation of Nitrocellulose. Anal. Lett. 2021, 54, 1929–1942. [Google Scholar] [CrossRef]
  44. Qadri, S.; Ganoe, A.; Haik, Y. Removal and recovery of acridine orange from solutions by use of magnetic nanoparticles. J. Hazard. Mater. 2009, 169, 318–323. [Google Scholar] [CrossRef] [PubMed]
  45. Antil, B.; Olhan, S.; Vander Wal, R.L. Production of Graphitic Carbon from Renewable Lignocellulosic Biomass Source. Minerals 2025, 15, 262. [Google Scholar] [CrossRef]
  46. Vasileiadou, A.; Zoras, S.; Iordanidis, A. Bioenergy production from olive oil mill solid wastes and their blends with lignite: Thermal characterization, kinetics, thermodynamic analysis, and several scenarios for sustainable practices. Biomass Convers. Biorefinery 2023, 13, 5325–5338. [Google Scholar] [CrossRef]
  47. Andriantsiferana, C.; Mohamed, E.F.; Delmas, H. Photocatalytic degradation of an azo-dye on TiO2/activated carbon composite material. Environ. Technol. 2014, 35, 355–363. [Google Scholar] [CrossRef]
  48. Yuan, R.; Guan, R.; Shen, W.; Zheng, J. Photocatalytic degradation of methylene blue by a combination of TiO2 and activated carbon fibers. J. Colloid Interface Sci. 2005, 282, 87–91. [Google Scholar] [CrossRef]
  49. Bukhari, S.N.U.S.; Shah, A.A.; Liu, W.; Channa, I.A.; Chandio, A.D.; Chandio, I.A.; Ibupoto, Z.H. Activated carbon based TiO2 nanocomposites (TiO2@AC) used simultaneous adsorption and photocatalytic oxidation for the efficient removal of Rhodamine-B (Rh–B). Ceram. Int. 2024, 50, 41285–41298. [Google Scholar] [CrossRef]
  50. Thirumoolan, D.; Ragupathy, S.; Renukadevi, S.; Rajkumar, P.; Rai, R.S.; Saravana Kumar, R.M.; Hasan, I.; Durai, M.; Ahn, Y.-H. Influence of nickel doping and cotton stalk activated carbon loading on structural, optical, and photocatalytic properties of zinc oxide nanoparticles. J. Photochem. Photobiol. A Chem. 2024, 448, 115300. [Google Scholar] [CrossRef]
  51. Hamieh, M.; Tabaja, N.; Chawraba, K.; Hamie, Z.; Hammoud, M.; Tlais, S.; Hamieh, T.; Toufaily, J. Visible Light Photo-Fenton with Hybrid Activated Carbon and Metal Ferrites for Efficient Treatment of Methyl Orange (Azo Dye). Molecules 2025, 30, 1770. [Google Scholar] [CrossRef]
  52. Imraish, A.; Abu Thiab, T.; Al-Awaida, W.; Al-Ameer, H.J.; Bustanji, Y.; Hammad, H.; Alsharif, M.; ad Al-Hunaiti, A. In vitro anti-inflammatory and antioxidant activities of ZnFe2O4 and CrFe2O4 nanoparticles synthesized using Boswellia carteri resin. J. Food Biochem. 2021, 45, e13730. [Google Scholar] [CrossRef]
  53. Liu, Q.-X.; Zhou, Y.-R.; Wang, M.; Zhang, Q.; Ji, T.; Chen, T.-Y.; Yu, D.-C. Adsorption of methylene blue from aqueous solution onto viscose-based activated carbon fiber felts: Kinetics and equilibrium studies. Adsorpt. Sci. Technol. 2019, 37, 312–332. [Google Scholar] [CrossRef]
  54. Liu, P.; Huang, Y.; Zhang, X. Enhanced electromagnetic absorption properties of reduced graphene oxide-polypyrrole with NiFe2O4 particles prepared with simple hydrothermal method. Mater. Lett. 2014, 120, 143–146. [Google Scholar] [CrossRef]
  55. Parishani, M.; Nadafan, M.; Dehghani, Z.; Malekfar, R.; Khorrami, G.H.H. Optical and dielectric properties of NiFe2O4 nanoparticles under different synthesized temperature. Results Phys. 2017, 7, 3619–3623. [Google Scholar] [CrossRef]
  56. da Silva, M.T.P.; Pereira, L.C.J.; Silva, F.T.; Rocha, R.P.; Guerreiro, M.C.; Silva, A.M. Textural and photocatalytic characteristics of iron-cobalt based nanocomposites supported on SBA-15: Synergistic effect between Fe2+ and Fe0 on photoactivity. Microporous Mesoporous Mater. 2021, 310, 110582. [Google Scholar] [CrossRef]
  57. Luadthong, C.; Itthibenchapong, V.; Viriya-Empikul, N.; Faungnawakij, K.; Pavasant, P.; Tanthapanichakoon, W. Synthesis, structural characterization, and magnetic property of nanostructured ferrite spinel oxides (AFe2O4, A = Co, Ni and Zn). Mater. Chem. Phys. 2013, 143, 203–208. [Google Scholar] [CrossRef]
  58. Doğan, M.; Sabaz, P.; Bicil, Z.; Koçer Kizilduman, B.; Turhan, Y. Activated carbon synthesis from tangerine peel and its use in hydrogen storage. J. Energy Inst. 2020, 93, 2176–2185. [Google Scholar] [CrossRef]
  59. Baikousi, M.; Dimos, K.; Bourlinos, A.B.; Zbořil, R.; Papadas, I.; Deligiannakis, Y.; Karakassides, M.A. Surface decoration of carbon nanosheets with amino-functionalized organosilica nanoparticles. Appl. Surf. Sci. 2012, 258, 3703–3709. [Google Scholar] [CrossRef]
  60. Shendrik, R.; Kaneva, E.; Radomskaya, T.; Sharygin, I.; Marfin, A. Relationships between the structural, vibrational, and optical properties of microporous cancrinite. Crystals 2021, 11, 280. [Google Scholar] [CrossRef]
  61. Makofane, A.; Motaung, D.E.; Hintsho-Mbita, N.C. Photocatalytic degradation of methylene blue and sulfisoxazole from water using biosynthesized zinc ferrite nanoparticles. Ceram. Int. 2021, 47, 22615–22626. [Google Scholar] [CrossRef]
  62. Yousif, M.; Ibrahim, A.H.; Al-Rawi, S.S.; Majeed, A.; Iqbal, M.A.; Kashif, M.; Abidin, Z.U.; Arbaz, M.; Ali, S.; Hussain, S.A.; et al. Visible light assisted photooxidative facile degradation of azo dyes in water using a green method. RSC Adv. 2024, 14, 16138–16149. [Google Scholar] [CrossRef] [PubMed]
  63. Srivastava, M.; Singh, J.; Yashpal, M.; Gupta, D.K.; Mishra, R.K.; Tripathi, S.; Ojha, A.K. Synthesis of superparamagnetic bare Fe3O4 nanostructures and core/shell (Fe3O4/alginate) nanocomposites. Carbohydr. Polym. 2012, 89, 821–829. [Google Scholar] [CrossRef]
  64. Sriramulu, M.; Shukla, D.; Sumathi, S. Aegle marmelos leaves extract mediated synthesis of zinc ferrite: Antibacterial activity and drug delivery. Mater. Res. Express 2018, 5, 115404. [Google Scholar] [CrossRef]
  65. Bharathi, K.K.; Noor-A-Alam, M.; Vemuri, R.S.; Ramana, C.V. Correlation between microstructure, electrical and optical properties of nanocrystalline NiFe1.925Dy0.075O4 thin films. RSC Adv. 2012, 2, 941–948. [Google Scholar] [CrossRef]
  66. Tabaja, N.; Brouri, D.; Casale, S.; Zein, S.; Jaafar, M.; Selmane, M.; Toufaily, J.; Davidson, A.; Hamieh, T. Use of SBA-15 silica grains for engineering mixtures of oxides CoFe and NiFe for Advanced Oxidation Reactions under visible and NIR. Appl. Catal. B Environ. 2019, 253, 369–378. [Google Scholar] [CrossRef]
  67. Zinatloo-Ajabshir, S.; Salavati-Niasari, M. Preparation of magnetically retrievable CoFe2O4@SiO2@Dy2Ce2O7 nanocomposites as novel photocatalyst for highly efficient degradation of organic contaminants. Compos. Part B Eng. 2019, 174, 106930. [Google Scholar] [CrossRef]
  68. Saleh, T.S.; Badawi, A.K.; Salama, R.S.; Mostafa, M.M.M. Design and Development of Novel Composites Containing Nickel Ferrites Supported on Activated Carbon Derived from Agricultural Wastes and Its Application in Water Remediation. Materials 2023, 16, 2170. [Google Scholar] [CrossRef]
  69. Wei, Z.; Huang, S.; Zhang, X.; Lu, C.; He, Y. Hydrothermal synthesis and photo-Fenton degradation of magnetic MnFe2O4/rGO nanocomposites. J. Mater. Sci. Mater. Electron. 2020, 31, 5176–5186. [Google Scholar] [CrossRef]
  70. Varghese, D.; Joe Raja Ruban, M.; Joselene Suzan Jennifer, P.; AnnieCanisius, D.; Chakko, S.; Muthupandi, S.; Madhavan, J.; Victor Antony Raj, M. Comprehensive analysis of NiFe2O4/MWCNTs nanocomposite to degrade a healthcare waste—Tetracycline. RSC Adv. 2023, 13, 28339–28361. [Google Scholar] [CrossRef]
  71. Aladdin Jasim, N.; Esmail Ebrahim, S.; Ammar, S.H. Fabrication of ZnxMn1-xFe2O4 metal ferrites for boosted photocatalytic degradation of Rhodamine-B dye. Results Opt. 2023, 13, 100508. [Google Scholar] [CrossRef]
  72. Toor, A.T.; Verma, A.; Jotshi, C.K.; Bajpai, P.K.; Singh, V. Photocatalytic degradation of Direct Yellow 12 dye using UV/TiO2 in a shallow pond slurry reactor. Dye. Pigment. 2006, 68, 53–60. [Google Scholar] [CrossRef]
  73. Panda, N.; Sahoo, H.; Mohapatra, S. Decolourization of Methyl Orange using Fenton-like mesoporous Fe2O3-SiO2 composite. J. Hazard. Mater. 2011, 185, 359–365. [Google Scholar] [CrossRef] [PubMed]
  74. Yu, W.; Liu, R.; Xu, K.; Lin, H.; Yang, M. Roles of iron species and pH optimization on sewage sludge conditioning with Fenton’s reagent and lime. Water Res. 2016, 95, 124–133. [Google Scholar] [CrossRef] [PubMed]
  75. Sedlak, D.L. Factors Affecting the Yield of Oxidants from the Reaction of Nanoparticulate Zero-Valent Iron and Oxygen. Environ. Sci. Technol. 2008, 42, 1262–1267. [Google Scholar] [CrossRef]
  76. Yu, L.; Chen, J.; Liang, Z.; Xu, W.; Chen, L.; Ye, D. Degradation of phenol using Fe3O4-GO nanocomposite as a heterogeneous photo-Fenton catalyst. Sep. Purif. Technol. 2016, 171, 80–87. [Google Scholar] [CrossRef]
  77. Chen, C.; Liang, Y.; Zhang, W. ZnFe2O4/MWCNTs composite with enhanced photocatalytic activity under visible-light irradiation. J. Alloys Compd. 2010, 501, 168–172. [Google Scholar] [CrossRef]
  78. Gaikwad, P.V.; Kamble, R.J.; Mane-Gavade, S.J.; Sabale, S.R.; Kamble, P.D. Magneto-structural properties and photocatalytic performance of sol-gel synthesized cobalt substituted NiCu ferrites for degradation of methylene blue under sunlight. Phys. B Condens. Matter 2019, 554, 79–85. [Google Scholar] [CrossRef]
  79. Sun, J.; Lin, X.; Xie, J.; Zhang, Y.; Wang, Q.; Ying, Z. Facile synthesis of novel ternary g-C3N4/ferrite/biochar hybrid photocatalyst for efficient degradation of methylene blue under visible-light irradiation. Colloids Surfaces A Physicochem. Eng. Asp. 2020, 606, 125556. [Google Scholar] [CrossRef]
  80. Giri, A.; Kumar, H.; Verma, M. Comprehensive study on the photocatalytic degradation of methylene blue dye using undoped and Mg-doped cobalt spinel ferrite nanoparticles with emphasis on Mg0.5Co0.5Fe2O4. Transit. Met. Chem. 2025. [Google Scholar] [CrossRef]
  81. Quiton, K.G.N.; Lu, M.-C.; Huang, Y.-H. Synergistic degradation of Methylene Blue by novel Fe-Co bimetallic catalyst supported on waste silica in photo-Fenton-like system. Sustain. Environ. Res. 2022, 32, 21. [Google Scholar] [CrossRef]
  82. Vadivel, S.; Maruthamani, D.; Habibi-Yangjeh, A.; Paul, B.; Dhar, S.S.; Selvam, K. Facile synthesis of novel CaFe2O4/g-C3N4 nanocomposites for degradation of methylene blue under visible-light irradiation. J. Colloid Interface Sci. 2016, 480, 126–136. [Google Scholar] [CrossRef] [PubMed]
  83. Ferreira, M.E.C.; Soletti, L.d.S.; Bernardino, E.G.; Quesada, H.B.; Gasparotto, F.; Bergamasco, R.; Yamaguchi, N.U. Synergistic Mechanism of Photocatalysis and Photo-Fenton by Manganese Ferrite and Graphene Nanocomposite Supported on Wood Ash with Real Sunlight Irradiation. Catalysts 2022, 12, 745. [Google Scholar] [CrossRef]
Figure 1. Wide-angle XRD of AC and FeCr/AC.
Figure 1. Wide-angle XRD of AC and FeCr/AC.
Appliedchem 05 00015 g001
Figure 2. N2 adsorption–desorption isotherms of AC and FeCr/AC.
Figure 2. N2 adsorption–desorption isotherms of AC and FeCr/AC.
Appliedchem 05 00015 g002
Figure 3. FT-IR spectra of AC and FeCr/AC.
Figure 3. FT-IR spectra of AC and FeCr/AC.
Appliedchem 05 00015 g003
Figure 4. SEM images and EDX results for AC (a,c) and FeCr (b,d), respectively.
Figure 4. SEM images and EDX results for AC (a,c) and FeCr (b,d), respectively.
Appliedchem 05 00015 g004
Figure 5. UV-Vis absorption spectra of AC and FeCr/AC (a), and Tauc plot of FeCr/AC (b).
Figure 5. UV-Vis absorption spectra of AC and FeCr/AC (a), and Tauc plot of FeCr/AC (b).
Appliedchem 05 00015 g005
Figure 6. MB removal efficiency under various experimental conditions ([MB] = 20 mg L−1, 250 ppm FeCr/AC, 0.25 mL H2O2, and pH = 5.47).
Figure 6. MB removal efficiency under various experimental conditions ([MB] = 20 mg L−1, 250 ppm FeCr/AC, 0.25 mL H2O2, and pH = 5.47).
Appliedchem 05 00015 g006
Figure 7. Effect of FeCr/AC mass on the degradation of MB.
Figure 7. Effect of FeCr/AC mass on the degradation of MB.
Appliedchem 05 00015 g007
Figure 8. Effect of initial MB concentration on its degradation efficiency using FeCr/AC.
Figure 8. Effect of initial MB concentration on its degradation efficiency using FeCr/AC.
Appliedchem 05 00015 g008
Figure 9. Effect of pH on the degradation efficiency of MB using FeCr/AC.
Figure 9. Effect of pH on the degradation efficiency of MB using FeCr/AC.
Appliedchem 05 00015 g009
Figure 10. Influence of H2O2 dosage on the degradation performance of MB.
Figure 10. Influence of H2O2 dosage on the degradation performance of MB.
Appliedchem 05 00015 g010
Figure 11. Degradation of MO and TCH.
Figure 11. Degradation of MO and TCH.
Appliedchem 05 00015 g011
Table 1. Textural characteristics of AC and FeCr/AC measured by N2 sorption at 77K.
Table 1. Textural characteristics of AC and FeCr/AC measured by N2 sorption at 77K.
SampleSBET (m2 g−1)Vp (cm3 g−1)Dpore (nm)Vµp (cm3 g−1)
Pure AC11480.63.20.055
FeCr/AC2220.329.70.025
Table 2. Comparison of band gap energies of different metal ferrites.
Table 2. Comparison of band gap energies of different metal ferrites.
Ferrite CatalystsBand Gap (eV)Ref.
NiFe2O4/SBA-152.09[66]
CoFe2O4@SiO2@Dy2Ce2O73.25[67]
NiFe2O4/AC (10NFAC)2.01[68]
MnFe2O4/rGO2.13[69]
CrFe2O4/SBA-152.7[51]
NiFe2O4/MWCNTs2.32[70]
CrFe2O4/AC1.9This study
Table 3. First-order kinetic parameters for MB degradation using FeCr/AC under different initial concentrations.
Table 3. First-order kinetic parameters for MB degradation using FeCr/AC under different initial concentrations.
Concentrations (ppm)k (min−1)R2t1/2 (min)
100.02970.978523.33
200.02580.973926.9
1000.00770.987389.78
Table 4. First-order kinetic parameters for MO and TCH.
Table 4. First-order kinetic parameters for MO and TCH.
PollutantFirst-Order Kinetic
k (min−1)R2t1/2 (min)
MO0.01150.998860.27
TCH0.02250.973130.8
Table 5. Comparison of the performance of various ferrite photocatalysts for MB degradation under different experimental conditions.
Table 5. Comparison of the performance of various ferrite photocatalysts for MB degradation under different experimental conditions.
PhotocatalystCatalyst Concentration (g L−1)[MB]0
(ppm)
pHLight SourceMB Degradation (%)Time (min)Ref.
NiFe2O4/MWCNTs110-Metal halide (400 W)99360[77]
Ni0.3Co0.2Cu0.5Fe2O4110-Visible (125 W)75120[78]
Ni0.1Co0.9Fe2O4/g-C3N4/biochar0.520-Xenon lamp (500 W)96.7120[79]
Mg0.5Co0.5Fe2O422510.5UV lamp (40 W)95.76120[80]
FeCo/SiO21203UV lamp (λ = 365 nm, 36 W)10060[81]
CaFe2O4/g-C3N4210 Tungsten halogen lamp (150 W)94120[82]
MnFe2O4-G@WA0.25107Sunlight (90–140 W/m2)94120[83]
FeCr/AC0.25206.47LED (25 W)97.56120This 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

Hamieh, M.; Al Khawand, S.; Tabaja, N.; Chawraba, K.; Hammoud, M.; Tlais, S.; Hamieh, T.; Toufaily, J. Chromium Ferrite Supported on Activated Carbon from Olive Mill Solid Waste for the Photo-Fenton Degradation of Pollutants from Wastewater Using LED Irradiation. AppliedChem 2025, 5, 15. https://doi.org/10.3390/appliedchem5030015

AMA Style

Hamieh M, Al Khawand S, Tabaja N, Chawraba K, Hammoud M, Tlais S, Hamieh T, Toufaily J. Chromium Ferrite Supported on Activated Carbon from Olive Mill Solid Waste for the Photo-Fenton Degradation of Pollutants from Wastewater Using LED Irradiation. AppliedChem. 2025; 5(3):15. https://doi.org/10.3390/appliedchem5030015

Chicago/Turabian Style

Hamieh, Malak, Sireen Al Khawand, Nabil Tabaja, Khaled Chawraba, Mohammad Hammoud, Sami Tlais, Tayssir Hamieh, and Joumana Toufaily. 2025. "Chromium Ferrite Supported on Activated Carbon from Olive Mill Solid Waste for the Photo-Fenton Degradation of Pollutants from Wastewater Using LED Irradiation" AppliedChem 5, no. 3: 15. https://doi.org/10.3390/appliedchem5030015

APA Style

Hamieh, M., Al Khawand, S., Tabaja, N., Chawraba, K., Hammoud, M., Tlais, S., Hamieh, T., & Toufaily, J. (2025). Chromium Ferrite Supported on Activated Carbon from Olive Mill Solid Waste for the Photo-Fenton Degradation of Pollutants from Wastewater Using LED Irradiation. AppliedChem, 5(3), 15. https://doi.org/10.3390/appliedchem5030015

Article Metrics

Back to TopTop