Translate this page into:
Advanced nanocomposite catalysts for waste management: Fe₃O₄-doped ZnO for dye degradation
* Corresponding author: E-mail address: a.mostafa@qu.edu.sa (Ayman M. Mostafa)
-
Received: ,
Accepted: ,
Abstract
This study investigates the synthesis and characterization of Fe3O4/ZnO nanocomposites via pulsed laser ablation (PLAL) in liquid media, focusing on their structural, optical, and photocatalytic properties. Zinc oxide (ZnO) nanoparticles were initially synthesized by ablating a zinc target in distilled water, followed by embedding with Fe3O4 nanoparticles through subsequent ablation of an iron target in the ZnO colloidal solution. X-ray diffraction showed the successful formation of a nanocomposite structure, revealing enhanced crystallinity and lattice parameters influenced by Fe3O4 doping. Optical studies demonstrated a redshift in the bandgap energy and increased Urbach energy, indicating modifications in electronic structure and surface defects due to doping. Photocatalytic degradation experiments using methyl orange (MO) dye under UV irradiation showed that the Fe3O4/ZnO nanocomposite exhibited superior photocatalytic efficiency (88.1%) compared to pure ZnO (77.7%) and Fe3O4 (60.2%), attributed to improved charge carrier separation and light absorption. These findings highlight the potential of Fe3O4/ZnO nanocomposites for environmental remediation applications.
Keywords
Laser ablation
Nanocomposite
Nanoparticles
Optical properties
Water treatment

1. Introduction
Zinc oxide (ZnO) is a semiconductor material from group II-VI with a wide energy bandgap of 3.37 eV and a notable exciton binding energy of 60 meV. It demonstrates minimal expense and low toxicity. ZnO can display crystalline structures characterized by four oxygen atoms in tetrahedral coordination surrounding the zinc atom [1]. Its characteristics have garnered global interest from researchers seeking to develop novel ZnO structures for various applications [2]. Moreover, in light of water constraints and population growth, ZnO is regarded as an effective material for removing refractory contaminants through heterogeneous photocatalysis, which is classified as an advanced oxidative reaction, wherein the generation of potent oxidant species facilitates the destruction of contaminant compounds [3]. The acquisition of ZnO is significantly affected by the conditions employed in the reaction medium. Factors such as the reaction temperature and precursor qualities can dramatically influence this oxide’s microstructural, optical, and morphological attributes [4-7].
The most common synthesis methods reported in the literature are chemical synthetic methods such as sol-gel, chemical precipitation, hydrothermal, and solvothermal methods. Every one of these choices has unique benefits and drawbacks [8-10]. However, the laser ablation was recently used in different fields as analytical tools [11-13] or synthetic tools in different types of medium in the film of colloidal nanoparticles or thin film. Liquid media technology’s pulsed laser ablation (PLAL) may create various nanoscale structures, sizes, and shapes. With this method, a high-power laser up to an extreme level is collimated and focused on the upper surface of the submerged solid for a predetermined time, resulting in a laser ablation process in a liquid medium. When the ablation threshold is exceeded, the material at the focus point first melts and then vaporizes. After that, the liquid turned into a gas after evaporating. The liquid’s vaporous emissions affect the solid, melting and breaking into nanoscale droplets that are then super-cooled by the liquid around them [14-20]. This approach is straightforward and unrestricted since it can create nanoparticles without surfactants or counter-ions. Because there are no chemical interactions, the nanoparticle’s surface is free of unnecessary ions or molecules. Using a multi-beam approach, the original laser beam’s nanostructured droplets can be broken up into smaller droplets by succeeding pulses. In addition, this technique allows the generation of nanostructured materials that are smaller and more uniform in size than others. Stable colloids, including nanostructured metal particles, can be produced by PLAL, eliminating the need for dispersants or surface-reactive chemicals. By using this technique, a certain kind of nanostructured material can be created inside a liquid solution. Because of this, novel nanocomposite structures with fascinating energy band gaps can be made [21-25].
In the decoration or embedding of the semiconductors, the wide bandgap semiconductors demonstrate promise for different applications, which could be modified with another nanostructured material, such as photocatalytic degradation based mainly on the optical characteristics of the modified nanocomposite structure [26,27]. Diverse approaches for synthesizing ZnO have been employed to enhance the progression of innovative technologies. This scenario involves metal doping, wherein an external metal ion is deliberately utilized to substitute the Zn2+ ion within the oxide host lattice [28-31]. Doping is an efficient technique for altering ZnO’s electro-optical properties, directing these innovative materials toward certain applications. The concentration, distribution, and chemical composition of dopants are essential factors that can directly affect the properties of ZnO. Therefore, transition metals are great candidates for introduction into the ZnO host lattice, which was created using the environmentally benign approach of PLAL in liquid media to achieve the desired results. It is common knowledge that doping with transition metal ions causes the introduction of various defects in the structure of the semiconductor, which in turn causes changes in the optical properties of ZnO, such as Fe3O4, which were created using the same method [32-36]. Consequently, different characterization techniques based on optical and structural investigation directly associated with zinc oxide and its enrichment with iron oxide nanoparticles are applied as a water treatment against MO dye.
2. Materials and Methods
2.1. Materials
Zinc sticks and iron rods, which had a purity of more than 99.99%, were purchased from Sigma-Aldrich, Germany.
2.2. Preparation method
ZnO NPs and their embedding with Fe3O4 NPs were prepared via the PLAL technique based on the following procedure: Zn target was used as a source of ZnO NPs by inserting this target at the bottom of the vial containing ultra-pure water. The target was ablated by the pulsed laser beam (1064 nm, Nd: YAG laser, Continuum, Surelite I, USA) that focused on the surface of the target by the plano-convex lens. The second step was similar to the first by changing the metal target to an Fe rod to generate Fe3O4 NPs. The last is related to creating ZnO NPs embedding via Fe3O4 NPs via ablation of Fe rod in the ZnO NPs solution that is produced from the first step to generate Fe3O4 NPs and decorate these NPs, the host material of ZnO NPs, as outlined in Figure 1. All appropriate parameters of the synthetic materials were mentioned in Table 1.

- Schematic diagram of the preparation of Fe3O4 embedding on ZnO NPs via PLAL.
| Sample code | 1st step | 2nd step | Description of the resultant product |
|---|---|---|---|
| ZnO | Ablation of Zn for 15 min in ultra-pure water | - | Colloidal ZnO NPs |
| Fe3O4 | - | Colloidal Fe3O4 NPs | |
| Fe3O4/ZnO | Ablation of Fe for 15 min in the prepared colloidal ZnO solution from 1st step | Embedding of ZnO NPs with Fe3O4 (Dopant is Fe3O4, and the matrix is ZnO) |
2.3. Photocatalytic degradation
The research was employed to decompose dye in an aqueous solution. The prepared solutions (dye and photocatalytic materials) were irradiated with a commercial lamp (160 W) functioning as a UV radiation source devoid of a bulb. The absorption in darkness for 30 min was recorded before initiating the photocatalytic test. The obtained aliquots were centrifuged and examined using a characteristic tool related to optical properties to study the photocatalytic efficiency was assessed utilizing Eq. (1) [37,38].
Where Co and Ct represent the starting up dye concentration and its concentration at any time, respectively, the prepared nanocomposite was added into photocatalytic systems, and the degradation rate was determined in each case.
2.4. Characterization techniques
UV-VIS 950 by PERKIN ELMER Spectrophotometer and X-ray diffractometer (Shimadzu XRD 6000, Japan).
3. Results and Discussion
3.1. Structural study
Figure 2(a) represents the diffractogram of the synthesized samples of ZnO, Fe3O4, and Fe3O4-doped ZnO created by ablation of a Zn target in ultra-pure water, laser ablation of a Fe target in ultra-pure water, and laser ablation of a Fe target in ultra-pure water containing a ZnO colloidal solution, respectively. X-ray diffraction (XRD) analysis analyzes these samples’ purity, crystallinity, and lattice structure. Ablation of the Zn target in ultra-pure water produces a diffractogram spectrum with 2θ values of 32.50°, 35.16°, 36.97°, 46.16°, 57.22°, 63.41°, 67.02°, 68.41°, and 69.70°, respectively. These values correspond to the Miller indices of (100), (002), (101), (102), (110), (103), (200), (112), and (201) crystal planes. The measured diffraction patterns suggest a hexagonal wurtzite ZnO structure with an average crystallite size of 44.52 nm. The hexagonal wurtzite structure of ZnO matches the reference card number (JCPDS 01-079-0205) [39]. The ZnO peaks suggest a crystalline structure. No extra ZnO XRD peaks indicate high purity. Eq. (2) is the Debye-Scherrer Equation, which was used to determine the average crystalline size of the generated samples. The peak at an angle of 36.97° indexed with (hkl) value (101) has a d-spacing of 2.46 Å and a Full width at half maximum (FWHM) of 0.21264. In the instance of laser ablation of a Fe target in ultra-pure water, the diffractogram spectrum reveals 2θ values of 18.87, 30.78, 36.02, 37.10, 43.58, 53.82, 57.43, and 62.98 degrees, which correspond to the (111), (220), (311), (222), (400), (422), (333), and (400) crystal planes, respectively. The observed diffraction patterns may be distinctly ascribed to the specified FCC structure of Fe3O4, characterized by the (227) space group and oriented along the (311) direction, with an average crystallite size of 77.11 nm. The appearance of the crystallite peaks corresponds to the reference card number (JCPDS 01-074-0748) [40], indicating a face-centered cubic magnetite (Fe3O4) structure. The peak at an angle of 36.11° indexed with (hkl) value (311) has a d-spacing of 2.46 Å and a FWHM of 0.17239. Furthermore, in the instance of laser ablation of the Fe target in ultra-pure water containing a ZnO colloidal solution, the simultaneous presence of mixed diffraction peaks verified the successful creation of the nanocomposite structure, including ZnO and Fe3O4.

- (a) XRD diffractogram of ZnO NPs, Fe3O4 NPs, and their nanocomposite Fe3O4/ZnO, and (b) Williamson-Hall plot for ZnO NPs and (c) Fe3O4/ZnO NPs.
Where k=0.9 is the shape factor, λ is the wavelength of the CuKα source (0.154 nm), β denotes FWHM, and “θ” is the Bragg’s angle [41]. Furthermore, in this embedding process, it was expected to find a change in the defect structure, which can be investigated via dislocation density (δ) [42,43], which was calculated using Eq. (3).
The insertion of Fe3O4 concentration into ZnO also increased the dislocation density. The density of dislocations in pure ZnO is 5.04 x 1014, while the density of dislocations in ZnO doped with Fe3O4 is 4.19 x 1014. The XRD graph proves this dislocation is present because the Fe content displays additional minor peaks. The Fe content results in more system defects than pure ZnO. So, incorporating Fe3O4 causes the movement of ZnO atoms toward grain boundaries. The dislocation density may decrease due to the dopant migration from the crystallite to the grain boundary. Furthermore, the interconnection between lattice parameters can be used to indicate the structure of the lattice. For the matrix structure of ZnO, the lattice constants (a and c) are estimated from eq. (4) [44], which were used to inter-planar spacing (dhkl) for a given plane with Miller indices (hkl) and the lattice parameters a (=b) and c as shown in eq. (4) [45] and tabulated in Table 2.
| Materials | FWHM | D, nm | δ x1014 | a=b, Ao | c, Ao | l, Ao | μ | DCorrected, nm |
|---|---|---|---|---|---|---|---|---|
| ZnO | 0.21264 | 44.52 | 5.04 | 2.806 | 4.860 | 4.901 | 0.0034 | 65.49 |
| Fe3O4/ZnO | 0.19379 | 48.85 | 4.19 | 2.808 | 4.863 | 4.905 | 0.0014 | 259.74 |
According to Table 2, the lattice parameters show that the increasing magnitudes of the “a” and “c” values with doping provide further confirmation of the fact that the XRD data had previously suggested that the lattice parameters rise with the concentration of the dopant (Figure 2). Consequently, the magnitudes of “a” and “c” experience a sudden increase if there is a presence of Fe3O4 concentration in ZnO. This increase in the lattice constants by embedding ZnO NPs with Fe3O4 NPs further supports the prior claim that the metallic oxide of Fe3O4 with a higher radius than that of Zn+2 enters the ZnO matrix. This is supported by the crystallite measurement of Fe3O4, which is observed in the material. In addition, the Williamson-Hall plot, which is indicated in eq. 5, can be utilized to analyze the lattice strain and the real particle size [44].
Where is the actual crystallite size of the prepared sample, while is the corrected strain, which could be measured by studying the relation of (β cos θ)/λ against (sin θ)/λ that known as the Williamson-Hall plot as mentioned in Figures 2(b) and (c). From this figure, the slope value is 0.0034 ± 0.0006.655, representing the corrected strain. Additionally, the crystallite size could be analyzed from the intercept, which showed 65.49 ± 0.00198 nm, whereas for the Fe3O4/ZnO case, the corrected strain is 0.0039 ± 0.00065, and the crystallite size is 259.74 ± 0.00651 nm. It can be said that the crystalline size calculated using the Scherrer equation is smaller than that obtained through the Williamson−Hall method; this change is attributed to the strain value and shows that strain is important and should, therefore, be considered when measuring the crystalline size.
3.2. Optical study
A crucial characteristic of nanocomposites in water purification technology is their optical characteristics. The optical properties of ZnO, Fe3O4, and Fe3O4-doped ZnO, synthesized through pulsed laser ablation of a Zn target in ultra-pure water, laser ablation of a Fe target in ultra-pure water, and laser ablation of a Fe target in ultra-pure water with a ZnO colloidal solution, were examined using UV-Vis absorption spectroscopy, as illustrated in Figure 3. The manufactured nanostructure’s optical properties result from each component’s combined effects, enabling our samples to absorb light across both the visible and ultraviolet bands. The absorbance of ZnO nanoparticles was assessed at wavelengths of 329, 258, and 190 nm. The initial peak pertained to the production of ZnO nanostructured material. In contrast, the subsequent peaks were associated with interband and intraband transitions between the valence and conduction bands, as well as phenomena linked to quantum confinement effects. The absorbance of Fe3O4 at the designated peaks was 325, 207, and below 190 nm, ascribed to analogous processes. The primary characteristic signal detected at 325 nm was due to electron excitation from occupied orbitals (t1u) to unoccupied anti-bonding states e*g and t*2g, thus confirming the presence of iron oxide nanoparticles in the suspension. The synthesis of the Fe3O4 and ZnO nanocomposite displayed absorbance peaks at 301 nm and below 190 nm, with the former signifying the emergence of a new nanoscale structure exhibiting quantum confinement and the latter of intra-band transitions. The absorbance spectra of Fe3O4-ZnO nanocomposites, relative to samples without a dopant metal oxide, show a significant enhancement in their physicochemical properties, evidenced by a marked shift of the absorbance peak from approximately 327 nm to 310 nm, influencing the optical and electronic characteristics of light absorption during photochemical interactions. The absorption characteristic pertains to electrical transitions induced by the presence of Fe3O4 within a hexagonal ZnO lattice. The absorption spectra indicate that doping levels have markedly enhanced absorption characteristics in both the visible and ultraviolet bands.

- (a) Absorbance spectra of the prepared ZnO NPs, Fe3O4 NPs, and their nanocomposite Fe3O4/ZnO, and (b-d) the fitting spectra of each of them.
The optical bandgap Eg was calculated using Eq. (6) and Eq. (7). This form was corrected and confirmed based on the Tauc relation as shown in Eq. (6) and Figure 4(a). From Figure 4(a), Eg values are 3.78, 3.76, and 4.06 eV for ZnO, Fe3O4, and Fe3O4/ZnO, respectively. An increase in Eg is related to the quantum confinement phenomena [46].

- (a) typical relation of (ahu)2 vs hu, (b-c) lnα versus photon energy (hυ), and (d) the tuning diagram of Eu and Eg, and (e) schematic diagram of transition between valence band and conduction band concerning Urbach and phonon energy for ZnO NPs, Fe3O4 NPs, and their nanocomposite Fe3O4/ZnO.
Besides, the localized states with microstructural lattice instabilities and crystal defects have band tails that match the Urbach energy or band tail energy (EU), which represents a very important study to know the reason for the modification in the photocatalytic degradation based on the embedding of ZnO with Fe3O4 NPs and can be calculated as in Eq. (8) [47].
where α0 is invariant. The Urbach relation explains crystalline materials’ absorption tail. Structural and synthesis disorders, point defects, deformations, and band gap pollutants cause the banded tail. Increased structural faults raise Urbach energy. Doping creates localized band states and strain potential in ZnO, changing its structural and electrical characteristics. The analyzed composition of the produced nanocomposite containing ZnO NPs and Fe3O4 NPs, as depicted in Figures 4(b) and (c), may be ascertained by extrapolating a linear graph to the zero X-axis value, which represents the Urbach plot against the natural logarithm of the incident photon energy [48,49]. The observations demonstrate that incorporating dopants increases Urbach energy, implying a heightened level of disorder in the system and, subsequently, surface defects in the ZnO matrix. So, Fe3O4 dopants deform lattices, as seen in Figure 4(d) for Eg. Figure 4(e) shows the energy level diagram’s Urbach area, emphasizing dopant inclusion at the conduction and valence band borders and Fe3O4/ZnO nanocomposite Eg and EU fluctuations.
The linear refractive index is a critical optical property to measure in nanocomposite materials, particularly when these materials are employed in photocatalytic degradation processes. The refractive index provides valuable insights into how light interacts with the material, which is fundamental to the efficiency and effectiveness of photocatalysis. In photocatalytic degradation, the primary mechanism involves the material’s absorption of light (typically UV or visible light), leading to the generation of e-h pairs. These charge carriers then participate in redox reactions that degrade organic pollutants or other target substances. The refractive index directly influences the material’s ability to absorb and utilize light, as it determines how light propagates through the material, how much is reflected at the surface, and how much is scattered or transmitted. A well-optimized refractive index can enhance light trapping within the material, ensuring that more incident light is absorbed rather than reflected or lost. This is particularly important in nanocomposites, where the integration of nanoparticles into a matrix can alter the optical properties significantly. Moreover, the refractive index is closely tied to the electronic structure and bandgap of the material, which are key factors in determining its photocatalytic activity. By measuring the refractive index, researchers can better understand how the nanocomposite interacts with photons of specific energies. This information is crucial for bandgap engineering, where the goal is to tailor the material to absorb light in the desired wavelength range, such as visible light for solar-driven photocatalysis. For instance, a material with a refractive index that facilitates strong absorption in the visible spectrum can harness a larger portion of solar energy, making it more efficient for environmental applications like water purification or air remediation. In addition to absorption, the refractive index affects light scattering and reflection at the surface and within the material. In photocatalytic systems, minimizing reflection losses and maximizing light penetration are essential for ensuring the material can utilize as much incident light as possible. A mismatch in refractive indices between the nanocomposite and its surrounding medium (e.g., air or water) can lead to significant reflection losses, reducing the amount of light available for photocatalysis. Researchers can design materials with reduced reflection and enhanced light absorption by measuring and optimizing the refractive index, thereby improving the overall photocatalytic performance [50]. Furthermore, the refractive index can provide insights into the dispersion of nanoparticles within the composite matrix, as agglomeration or poor distribution can lead to inhomogeneous optical properties and reduced efficiency. In summary, measuring the linear refractive index in nanocomposite materials is paramount for photocatalytic degradation applications. It helps understand and optimize light-matter interactions and plays a key role in bandgap engineering, minimizing optical losses, and enhancing the overall efficiency of the photocatalytic process. By carefully tailoring the refractive index, researchers can develop advanced nanocomposites that maximize light absorption, improve charge carrier generation, and lead to more effective and sustainable photocatalytic systems for environmental remediation and other applications. There are many models, including Moss (nM), Reddy-Ahammed (nR-Ah), Ravindra (nR), Herve and Vandamme (nHV), Kumar-Singh (nKS), and Annani et al. [51] (nA), demonstrated that the inverse relationship between Eg and the linear refractive index is as the following Eq. (9) [52,53].
The refractive indices were the same across all models; nM, nR-Ah, nR, nHV, nKS, and nA were computed for the six corresponding models. As packing density diminishes, the material’s reflectance decreases while its transparency improves with a reduction in particle size. Figure 5(a) illustrates the variation of refractive indices according to each model. At the same time, Figure 5(b) discusses the alteration of the average refractive index of the synthesized nanostructures based on all specified parameters, indicating a reduction in the nanocomposite formation due to low packing density, resulting in a decrease in the material’s reflectance.

- (a) Refractive indices based on different models and (b) the average refractive index for ZnO NPs, Fe3O4 NPs, and their nanocomposite.
3.3. Photocatalytic study
3.3.1. Photocatalytic degradation
The photocatalytic destruction of metal oxide (MO), a model pollutant [54], under UV light irradiation in an aqueous solution is employed to assess the photocatalytic effectiveness of Fe3O4/ZnO composites. The optimization process is initially performed to identify the ideal synthesis conditions for the Fe3O4/ZnO nanocomposite. The photocatalytic degradation of ZnO, Fe3O4, and Fe3O4-doped ZnO, synthesized through pulsed laser ablation of a Zn target in ultra-pure water, laser ablation of a Fe target in ultra-pure water, and laser ablation of a Fe target in ultra-pure water containing a ZnO colloidal solution, respectively, was examined, as illustrated in Figure 6 [55]. This figure presents the absorption spectra of 60 mg of the synthesized photocatalyst, evaluated for photocatalytic degradation of MO dye, with the absorption spectrum displayed as a function of time. This figure indicates no substantial variation in the photocatalytic degradation percentage of MO dye in the absence of a catalyst under UV light. The peak photocatalytic efficiency was recorded at 77.7%, 60.2%, and 88.1% for ZnO, Fe3O4, and Fe3O4-doped ZnO, respectively.

- (a) The behavior degradation spectra process of MO dye, (b) The intensity change of the main characteristic peak of MO dye at 478 nm, and (c) the degradation efficiency versus reaction time of ZnO NPs, Fe3O4 NPs, and their nanocomposite Fe3O4/ZnO.
3.3.2. Optimum condition of photocatalytic degradation
The ZnO and Fe3O4 NP-based samples’ photocatalytic efficiency is tested by degrading MO in an aqueous solution exposed to UV light. The initial step is to optimize Fe3O4/ZnO nanocomposite synthesis parameters, such as UV-irradiation period, catalyst dose, and dye concentration (Figure 7). According to Figures 7(a) and (b), dark conditions were upheld for the adsorption and sonication treatments for the initial 30 min. The rate of degradation escalated with the duration of UV irradiation. The photodegradation of MO utilizing the examined samples under sonication yielded optimal activity. The maximum absorbance of MO, corresponding to the absorption peak at 478 nm, diminished steadily with extended UV irradiation duration. As indicated by this figure, the catalyst dosages for the photocatalytic degradation of dyes are a critical parameter. The dosage was adjusted from 10 mg to 30 mg for a 60 mg/L MO solution to examine the impact of catalyst weight. It was noted that photodegradation escalates with an increase in the quantity of catalyst. Nevertheless, when the catalyst quantity surpasses the optimal level (25 mg), the efficacy of photo-degradation diminishes. The increase in available adsorption and catalytic sites on the photocatalyst was responsible for this; however, at elevated concentrations, the turbidity of the MO suspension rises, diminishing UV light penetration, which in turn amplifies the scattering effect and reduces the photo-degradation rate. Furthermore, in Figure 7(c), the impact of initial MO concentration on photo-degradation efficiency was examined by adjusting the concentration to 60 mg while maintaining the photocatalyst quantity at 25 mg/10 ml. The photo-degradation effectiveness of MO diminished as the initial dye concentration increased. As the dye concentration increases, photons are obstructed before reaching the catalyst surface, diminishing the photon absorption. Consequently, the dye molecules may absorb a substantial quantity of photons instead of the catalyst, diminishing catalytic effectiveness [56-58].

- (a-b) Degradation efficiency and behavior of the rate catalyst at different catalyst dosages, and (c) degradation efficiency at varying the concentration of MO by using Fe3O4/ZnO as a photocatalyst.
3.3.3. Kinetic study
The kinetic behavior of methyl orange (MO) dye degradation and adsorption using ZnO NPs, Fe₃O₄ NPs, and Fe₃O₄/ZnO nanocomposite was systematically investigated. Figure 8 provides a detailed, multi-panel representation of both the dynamic degradation profile and the kinetic modeling for photocatalysis and adsorption processes. Figure 8(a) illustrates the temporal evolution of MO concentration under UV irradiation in the presence of the three catalysts. The Fe₃O₄/ZnO nanocomposite demonstrates a significantly enhanced degradation rate compared to the individual ZnO and Fe₃O₄ nanoparticles. This superior performance is attributed to the synergistic effect between ZnO and Fe₃O₄, which likely improves charge separation and increases the number of active sites for photocatalytic reactions [59].

- (a) Dynamic profile of the catalytic degradation of MO dye, and kinetic study of (b-c) MO-photocatalytic degradation: (b) pseudo 1st order, (c) pseudo 2nd order, and (d-f) MO-adsorption study via (d) pseudo 1st order, (e) pseudo 2nd order, (f) intra-particle-diffusion (e) by ZnO NPs, Fe3O4 NPs, and their nanocomposite Fe3O4/ZnO.
To elucidate the underlying reaction mechanisms, the photocatalytic degradation data were fitted to both pseudo-first-order (Figure 8b) and pseudo-second-order (Figure 8c) kinetic models. In Pseudo-First-Order Model (Figure 8b), the linearity of the plots for all catalysts indicates that the photocatalytic degradation of MO can be approximated by a pseudo-first-order reaction, as described by the Langmuir–Hinshelwood mechanism. The Fe₃O₄/ZnO nanocomposite exhibits the highest rate constant, confirming its superior catalytic activity. In other words, in the pseudo-second-order model (Figure 8c), the fitting to the pseudo-second-order model was also performed. While some correlation is observed, the regression coefficients (R2 values) are generally lower than those for the first-order model, suggesting that the degradation process is better described by the pseudo-first-order kinetics for all samples. The photocatalytic degradation of MO dye demonstrated a pseudo-first-order response, characterized by the kinetics of the photocatalyst as ln(C/Co) = kt. After fitting the specified model, the k and R2 values are obtained and displayed in Table 3. The tabulated data demonstrate that the rate constant (k), R2 value, and degradation percentage are more favorable for ZnO/Fe3O4 than for the other samples analyzed [60-63].
| Photocatalytic sample | The rate constant, K1 | R2 | The rate constant, K2 | R2 | Degradation after 60 min, % |
|---|---|---|---|---|---|
| ZnO | 0.01577 | 0.982 | 0.0136 | 0.939 | 0.93 |
| Fe3O4 | 0.02658 | 0.989 | 0.0136 | 0.973 | 60.2 |
| ZnO/Fe3O4 | 0.03591 | 0.973 | 0.0134 | 0.914 | 88.1 |
From Figures 8(d-f), the adsorption of MO onto the catalyst surfaces prior to irradiation was analyzed using three kinetic models. Pseudo-first-order adsorption (Figure 8d), the experimental data for all samples fit reasonably well to the pseudo-first-order model, indicating that physisorption is a significant contributor to the initial dye removal. Pseudo-second-order adsorption (Figure 8e), the pseudo-second-order model provides an even better fit, especially for the Fe₃O₄/ZnO nanocomposite, as evidenced by higher R2 values. This suggests that chemisorption, involving valence forces through sharing or exchange of electrons, plays a dominant role in the adsorption process. Intra-particle diffusion (Figure 8f), the intra-particle diffusion plots reveal multi-linear behavior, indicating that the adsorption process is governed by more than one mechanism. The initial sharper region corresponds to surface adsorption, followed by a gradual region attributed to intraparticle diffusion. The Fe₃O₄/ZnO nanocomposite exhibits the highest slope and intercept, confirming its superior adsorption capacity and faster diffusion rate.
So, across all kinetic models, the Fe₃O₄/ZnO nanocomposite consistently outperforms the individual oxide nanoparticles. The enhanced photocatalytic and adsorption properties are likely due to improved electron–hole separation (reducing recombination), increased surface area and active sites, and synergistic effects between ZnO and Fe₃O₄ phases. The updated kinetic analysis, as visualized in Figure 8, clearly demonstrates that embedding Fe₃O₄ into ZnO nanoparticles significantly enhances both the photocatalytic degradation and adsorption of MO dye. The Fe₃O₄/ZnO nanocomposite follows pseudo-first-order kinetics for photocatalytic degradation and pseudo-second-order kinetics for adsorption, with intra-particle diffusion also playing a role. These findings confirm the potential of Fe₃O₄/ZnO nanocomposites as highly efficient catalysts for environmental remediation applications.
3.3.4. Reusability
The practical application of photocatalysts in environmental remediation critically depends on their long-term stability and reusability. To address these aspects, cyclic degradation tests were performed to evaluate the performance of the synthesized Fe3O4/ZnO nanocomposites over multiple photocatalytic cycles. After each degradation cycle of MO dye, the nanocomposite catalyst was separated, thoroughly washed with distilled water to remove residual dye and byproducts, and then reused under identical conditions for subsequent cycles. The photocatalytic efficiency was monitored by measuring the degradation percentage of MO after each cycle. The results demonstrated that Fe3O4/ZnO nanocomposites maintained a high level of photocatalytic activity over five consecutive cycles, with only a moderate decrease in efficiency. Specifically, the degradation efficiency remained above 74% after the fifth cycle (Figure 9), indicating excellent stability and reusability of the catalyst. The slight reduction in activity can be attributed to unavoidable catalyst loss during recovery and washing, as well as minor surface fouling or structural changes that may occur after repeated use. Nevertheless, the nanocomposite retained most of its catalytic performance, underscoring its robustness and suitability for practical wastewater treatment applications. These findings confirm that Fe3O4/ZnO nanocomposite not only exhibits superior initial photocatalytic activity but also demonstrate strong long-term operational stability and recyclability, making them a promising candidate for sustainable pollutant degradation in real-world scenarios.

- The degradation efficiency of the reusability of Fe3O4/ZnO up to five times.
3.3.5. Consumption and potential ecological toxicity
The synthesis of Fe₃O₄/ZnO nanocomposite via pulsed laser ablation in liquid, as employed in our study, offers several environmental advantages compared to conventional chemical methods. This technique does not require chemical surfactants or additional reagents, thereby minimizing chemical waste and potential secondary pollution. The process utilizes only high-purity metal targets and water, reducing the risk of introducing hazardous byproducts into the environment. Nevertheless, scaling up production would increase energy consumption and the demand for raw materials such as high-purity zinc and iron. We acknowledge that the environmental footprint of large-scale synthesis should be carefully assessed, including a life cycle analysis of energy and material inputs. Regarding ecological toxicity, Fe₃O₄/ZnO nanoparticles have demonstrated efficient removal of organic dyes and heavy metals from water, but concerns remain about their potential impact if released into natural ecosystems. Literature reports indicate that while these nanocomposites are generally considered to have low toxicity, their accumulation and persistence in aquatic environments could pose risks to microorganisms and higher trophic levels. To mitigate this, the magnetic properties of Fe₃O₄/ZnO enable efficient recovery from treated water, reducing the likelihood of nanoparticle release and supporting catalyst reuse. We have included these considerations in the revised discussion section and emphasized the need for further research into the long-term environmental safety and green synthesis approaches for Fe₃O₄/ZnO nanocomposites [64].
4. Conclusions
In contrast, the Fe₃O₄/ZnO nanocomposites synthesized via PLAL offer several distinct advantages. The PLAL technique is a surfactant-free, green synthesis method that eliminates the need for hazardous chemicals and post-synthesis purification, resulting in high-purity nanomaterials. The use of abundant and non-toxic starting materials (zinc and iron) further enhances the cost-effectiveness of this approach. While the initial capital investment for laser equipment may be higher than for some chemical methods, this cost is offset by the simplicity of the process, reduced chemical consumption, and minimal waste generation. From a scalability perspective, PLAL has traditionally been limited to laboratory-scale production. However, recent advances in laser technology and reactor design are making it increasingly feasible to scale up the process for larger batch or continuous production of nanomaterials. This positions PLAL-based nanocomposites as promising candidates for future industrial applications, provided further engineering optimization is achieved. Regarding long-term stability, the Fe₃O₄/ZnO nanocomposites demonstrate high crystallinity and structural integrity, as confirmed by XRD analysis. The absence of surfactants or organic residues on the nanoparticle surfaces is expected to enhance their chemical and photostability during repeated use. Additionally, the incorporation of Fe₃O₄ provides magnetic properties, which can facilitate easy separation and recovery of the catalyst from treated water, potentially improving reusability and operational efficiency. Nevertheless, some limitations remain. The throughput of PLAL is currently lower than that of conventional chemical synthesis routes, and the long-term field stability and regeneration performance of the nanocomposites require further investigation under real-world conditions. Additionally, while the magnetic separation feature is advantageous, it may not be necessary or cost-effective for all water treatment scenarios.
The embedding of Fe3O4 nanoparticles into ZnO via pulsed laser ablation successfully produced Fe3O4/ZnO nanocomposites with enhanced structural, optical, and photocatalytic properties. Incorporating Fe3O4 into the ZnO matrix resulted in lattice expansion, increased Urbach energy, and a tunable bandgap, collectively contributing to improved light absorption and charge carrier dynamics. The nanocomposite demonstrated superior photocatalytic performance in degrading MO dye under UV irradiation, achieving 88.1% degradation efficiency, compared to 77.7% for pure ZnO and 60.2% for Fe3O4. This enhancement is attributed to the synergistic effects of Fe3O4 doping, which reduced electron-hole recombination and increased active sites for catalysis. The study underscores the potential of Fe3O4/ZnO nanocomposites as efficient photocatalysts for environmental applications, particularly in wastewater treatment, while highlighting the versatility of pulsed laser ablation to design advanced nanomaterials. Future research could explore additional dopants or optimize synthesis parameters to enhance performance further.
In summary, the Fe₃O₄/ZnO nanocomposites produced by PLAL present a compelling combination of high photocatalytic efficiency, environmental friendliness, and potential for magnetic recovery. These advantages, balanced against current challenges in scalability and long-term deployment, highlight both the promise and the areas for further development of this approach in comparison to existing solutions in the field of water treatment and environmental remediation.
Acknowledgment
The authors extend their appreciation to the Deanship of Scientific Research and Libraries at Princess Nourah Bint Abdulrahman University for funding this research work through the Research Group project, Grant No. (RG-1445-0056).
CRediT authorship contribution statement
Nada S. Al-Kadhi, Ayman M. Mostafa: Conceptualization, Formal analysis, Investigation, Writing – review & editing. Eman A. Mwafy: Conceptualization, Methodology, investigation, Formal analysis, Data curation, Writing – review & editing. Ameenah N. Al-Ahmadi, Hoda A. Ahmed, and Fowzia S. Alamro: Conceptualization, Formal analysis, Investigation, Writing – original draft. Ghadah M. Al-Senani, Wafaa B. Elsharkawy, and Rawan Al-Faze: Conceptualization, Methodology, Formal analysis, Data curation, writing – original draft.
Declaration of competing interest
The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.
Data availability
No datasets were generated or analyzed during the current study.
Declaration of generative AI and AI-assisted technologies in the writing process
The authors confirm that there was no use of artificial intelligence (AI)-assisted technology for assisting in the writing or editing of the manuscript, and no images were manipulated using AI.
References
- Synthesis, antimicrobial and antioxidant activities of ZnO, PbO, and Al2O3 NPs obtained by calcining new coordination polymers. Results in Chemistry. 2024;11:101840. https://doi.org/10.1016/j.rechem.2024.101840
- [Google Scholar]
- High-efficiency photocatalyst based on Au nanoparticles loaded on defective ZnO nanorods. Vacuum. 2024;225:113236. https://doi.org/10.1016/j.vacuum.2024.113236
- [Google Scholar]
- Zinc oxide/carbon nanotubes nanocomposite: Synthesis, characterization and catalytic reduction of 4-nitrophenol via laser assistant method. Surfaces and Interfaces. 2021;26:101406. https://doi.org/10.1016/j.surfin.2021.101406
- [Google Scholar]
- Eco-friendly synthesis of Ag-ZnO Nanocomposite and its anti-bacterial activity, photocatalysis toward degradation of (CB) dye and removal wastewater pollution. International Journal of Nanoscience. 2023;22, 05:2350041. https://doi.org/10.1142/s0219581x23500412
- [Google Scholar]
- Fast one-pot laser-based fabrication of ZnO/TiO2-reduced graphene oxide nanocomposite for photocatalytic applications. Optics & Laser Technology. 2023;160:109105. https://doi.org/10.1016/j.optlastec.2022.109105
- [Google Scholar]
- Enhanced plasmonic photocatalysis of Au-decorated ZnO nanocomposites. Inorganics. 2023;11:157. https://doi.org/10.3390/inorganics11040157
- [Google Scholar]
- The role of the pulsed laser deposition in different growth atmospheres on the gas-sensing properties of ZnO films. Sensors and Actuators B: Chemical. 2023;382:133454. https://doi.org/10.1016/j.snb.2023.133454
- [Google Scholar]
- Synthesis and characterization of Sm3+ doped BaO-ZnO-LiF-B2O3 glass system for reddish-orange light generation with high color purity. Optics & Laser Technology. 2022;155:108359. https://doi.org/10.1016/j.optlastec.2022.108359
- [Google Scholar]
- Cu-doped ZnO-nanoparticles as a novel eco-friendly insecticide for controlling Spodoptera littoralis. Biocatalysis and Agricultural Biotechnology. 2023;52:102823. https://doi.org/10.1016/j.bcab.2023.102823
- [Google Scholar]
- Synthesis of Mesoporous Ru-ZnO@g-C3N4 nanoparticles and their photocatalytic activity for methylene blue degradation. Water. 2023;15:481. https://doi.org/10.3390/w15030481
- [Google Scholar]
- Synthesis and nonlinear optical properties of a novel indium phthalocyanine highly branched polymer. Polymers for Advanced Technologies. 2015;26:1014-1019. https://doi.org/10.1002/pat.3520
- [Google Scholar]
- Indium(III) phthalocyanine eka-conjugated polymer as high-performance optical limiter upon nanosecond laser irradiation. High Performance Polymers. 2016;28:651-659. https://doi.org/10.1177/0954008315593616
- [Google Scholar]
- Effect of single and double pulse laser-induced breakdown spectroscopy towards steel alloy in different gaseous media. Materials Chemistry and Physics. 2024;320:129443. https://doi.org/10.1016/j.matchemphys.2024.129443
- [Google Scholar]
- Preparation and study of nonlinear response of embedding ZnO nanoparticles in PVA thin film by pulsed laser ablation. Journal of Molecular Structure. 2021;1223:129007. https://doi.org/10.1016/j.molstruc.2020.129007
- [Google Scholar]
- Preparation and study of nonlinear response of embedding ZnO nanoparticles in PVA thin film by pulsed laser ablation. Journal of Molecular Structure. 2021;1223:129007. https://doi.org/10.1016/j.molstruc.2020.129007
- [Google Scholar]
- ZnO nanoparticles decorated carbon nanotubes via pulsed laser ablation method for degradation of methylene blue dyes. Colloids and Surfaces A: Physicochemical and Engineering Aspects. 2021;627:127204. https://doi.org/10.1016/j.colsurfa.2021.127204
- [Google Scholar]
- Nonlinearity enhancement of Multi-walled carbon nanotube decorated with ZnO nanoparticles prepared by laser assisted method. Optics & Laser Technology. 2022;155:108444. https://doi.org/10.1016/j.optlastec.2022.108444
- [Google Scholar]
- Influence of Urbach energy in enhanced photocatalytic activity of Cu doped ZnO nanoparticles. Optical Materials. 2022;127:112245. https://doi.org/10.1016/j.optmat.2022.112245
- [Google Scholar]
- Laser-assisted deposition of Ag on self-sourced growth ZnO nanorods as reusable photocatalysts for water purification. Inorganic Chemistry Communications. 2022;146:110065. https://doi.org/10.1016/j.inoche.2022.110065
- [Google Scholar]
- Tuning the morphology of Au/ZnO nanocomposite using pulsed laser ablation for anticancer applications. Arabian Journal for Science and Engineering. 2024;49:1063-1074. https://doi.org/10.1007/s13369-023-08061-8
- [Google Scholar]
- A new method for preparing ZnO/CNT nanocomposites with enhanced photocatalytic degradation of malachite green under visible light. Journal of Photochemistry and Photobiology A: Chemistry. 2020;389:112207. https://doi.org/10.1016/j.jphotochem.2019.112207
- [Google Scholar]
- Preparation and characterization of invert ZnO–B2O3 glasses and its shielding behavior towards gamma irradiation. Materials Chemistry and Physics. 2020;240:122129. https://doi.org/10.1016/j.matchemphys.2019.122129
- [Google Scholar]
- Enhancement of photocatalytic and Sonophotocatalytic degradation of 4-nitrophenol by ZnO/graphene oxide and ZnO/carbon nanotube nanocomposites. Journal of Photochemistry and Photobiology A: Chemistry. 2020;396:112507. https://doi.org/10.1016/j.jphotochem.2020.112507
- [Google Scholar]
- Synthesis of ZnO/CdO thin film for catalytic degradation of 4-nitrophenol. Journal of Molecular Structure. 2020;1221:128872. https://doi.org/10.1016/j.molstruc.2020.128872
- [Google Scholar]
- Synthesis of ZnO and Au@ZnO core/shell nano-catalysts by pulsed laser ablation in different liquid media. Journal of Materials Research and Technology. 2020;9:3241-3248. https://doi.org/10.1016/j.jmrt.2020.01.071
- [Google Scholar]
- Zeolitic imidazolate framework (ZIF-8) modified cellulose acetate NF membranes for potential water treatment application. Carbohydrate Polymers. 2023;299:120230. https://doi.org/10.1016/j.carbpol.2022.120230
- [Google Scholar]
- A review on hydrophobic electrospun nanofibers-based materials and membranes for water treatment: Challenges, outlook, and stability. Separation and Purification Technology. 2025;353:128370. https://doi.org/10.1016/j.seppur.2024.128370
- [Google Scholar]
- Composition/structural evolution and optical properties of ZnO/Zn nanoparticles by laser ablation in liquid media. The Journal of Physical Chemistry. B. 2005;109:18260-18266. https://doi.org/10.1021/jp052258n
- [Google Scholar]
- Violet luminescence emitted from Ag-nanocluster doped ZnO thin films grown on fused quartz substrates by pulsed laser deposition. Physica B: Condensed Matter. 2006;373:154-156. https://doi.org/10.1016/j.physb.2005.11.116
- [Google Scholar]
- Improved dye-sensitized solar cells with a ZnO-nanoflower photoanode. Applied Physics Letters. 2007;90 https://doi.org/10.1063/1.2751588
- [Google Scholar]
- Optical characterization of Sm3+ and Dy3+: ZnO–PbO–B2O3 glasses. Spectrochimica Acta Part A: Molecular and Biomolecular Spectroscopy. 2007;67:802-807. https://doi.org/10.1016/j.saa.2006.08.036
- [Google Scholar]
- Fe3O4-multiwalled carbon nanotubes-bentonite as adsorbent for removal of methylene blue from aqueous solutions. Chemosphere. 2023;316:137824. https://doi.org/10.1016/j.chemosphere.2023.137824
- [Google Scholar]
- Nanocomposite Fe3O4-MWCNTs based on femtosecond pulsed laser ablation for catalytic degradation. Diamond and Related Materials. 2023;140:110445. https://doi.org/10.1016/j.diamond.2023.110445
- [Google Scholar]
- Design and synthesize of folate decorated Fe3O4@Au-DEX-CP nano formulation for targeted drug delivery in colorectal cancer therapy: In vitro and in vivo studies. Journal of Drug Delivery Science and Technology. 2023;87:104798. https://doi.org/10.1016/j.jddst.2023.104798
- [Google Scholar]
- Evaluation of synthesized MWCNTs-COO@Fe3O4 nanocomposite based on laser-assisted method for highly-efficient clean water. Materials Chemistry and Physics. 2024;316:129093. https://doi.org/10.1016/j.matchemphys.2024.129093
- [Google Scholar]
- Ag@Fe3O4-coated U-shaped plastic optical fiber sensor for H2S detection. Sensors and Actuators B: Chemical. 2024;401:135054. https://doi.org/10.1016/j.snb.2023.135054
- [Google Scholar]
- Removal of methylene blue dye from aqueous solution using carbon nanotubes decorated by nickel oxide nanoparticles via pulsed laser ablation method. Radiation Physics and Chemistry. 2022;198:110268. https://doi.org/10.1016/j.radphyschem.2022.110268
- [Google Scholar]
- Nanoparticle-mediated remediation of wastewater contaminants: An inclusive analysis of glyphosate, Congo red and methyl orange. Journal of Molecular Structure. 2025;1321:140127. https://doi.org/10.1016/j.molstruc.2024.140127
- [Google Scholar]
- Solution combustion synthesis of Ni/NiO/ZnO nanocomposites for photodegradation of methylene blue under ultraviolet irradiation. Journal of Electronic Materials. 2018;47:2703-2709. https://doi.org/10.1007/s11664-018-6119-7
- [Google Scholar]
- Thiacalix[4]arene-functionalized magnetic xanthan gum (TC4As-XG@Fe3O4) as a hydrogel adsorbent for removal of dye and pesticide from water medium. Separation and Purification Technology. 2023;306:122700. https://doi.org/10.1016/j.seppur.2022.122700
- [Google Scholar]
- Effects of post-laser irradiation on the optical and structure properties of Al2O3 nanoparticles produced by laser ablation. Journal of Applied Physics. 2020;128:153104. https://doi.org/10.1063/5.0022554
- [Google Scholar]
- Physical, structural and spectroscopic studies of Al2O3-B2O3-Sm2O3 scintillating glass doped with heavy metal oxides. Journal of Luminescence. 2022;250:119093. https://doi.org/10.1016/j.jlumin.2022.119093
- [Google Scholar]
- Absorption spectrum fitting method (ASF), DASF method and structural studies of Li2O–SrO–B2O3–MnO quaternary glass system. Optical Materials. 2022;133:112911. https://doi.org/10.1016/j.optmat.2022.112911
- [Google Scholar]
- impact of radio frequency plasma power on the structure, crystallinity, dislocation density, and the energy band gap of ZnO nanostructure. ACS omega. 2021;6:31605-31614. https://doi.org/10.1021/acsomega.1c04105
- [Google Scholar]
- Structural feature, morphology and optical properties of g-C3N4 decorated CuO-Fe3O4 nano composite for electrocatalytic and photocatalytic applications. Diamond and Related Materials. 2024;147:111294. https://doi.org/10.1016/j.diamond.2024.111294
- [Google Scholar]
- UV-induced macromolecular and optical modifications in gelatin solid films with transition metal chlorides. Journal of Molecular Structure. 2019;1182:181-190. https://doi.org/10.1016/j.molstruc.2019.01.054
- [Google Scholar]
- Optical energy band gap investigation of the zinc carbide nanoparticles prepared by nanosecond pulsed laser ablation of zinc metals in hexane liquid medium. Optical Materials. 2024;152:115409. https://doi.org/10.1016/j.optmat.2024.115409
- [Google Scholar]
- Optical properties of Sm3+-doped CaF2 bismuth borate glasses. Journal of Luminescence. 2009;129:1314-1319. https://doi.org/10.1016/j.jlumin.2009.06.017
- [Google Scholar]
- Impact of Urbach energy on open-circuit voltage deficit of thin-film solar cells. Solar Energy Materials and Solar Cells. 2020;210:110502. https://doi.org/10.1016/j.solmat.2020.110502
- [Google Scholar]
- Facile synthesis of Cu2O nanoparticles using pulsed laser ablation method for optoelectronic applications. Colloids and Surfaces A: Physicochemical and Engineering Aspects. 2021;630:127562. https://doi.org/10.1016/j.colsurfa.2021.127562
- [Google Scholar]
- Model for calculating the refractive index of a III–V semiconductor. Computational Materials Science. 2008;41:570-575.
- [Google Scholar]
- Effect of reduced graphene oxide on the structural and optical properties of ZnO nanoparticles. Materials Letters. 2024;355:135465. https://doi.org/10.1016/j.matlet.2023.135465
- [Google Scholar]
- Ramification of Li-ion array on structural, optical and electrical properties of vanadium pentoxide for energy-efficient material. Electrochimica Acta. 2025;514:145659. https://doi.org/10.1016/j.electacta.2025.145659
- [Google Scholar]
- Defected Ag/Cu-MOF as a modifier of polyethersulfone membranes for enhancing permeability, antifouling properties and heavy metal and dye pollutant removal. Separation and Purification Technology. 2024;345:127336. https://doi.org/10.1016/j.seppur.2024.127336
- [Google Scholar]
- Mechanothermal synthesis of Ag/TiO2 for photocatalytic methyl orange degradation and hydrogen production. Process Safety and Environmental Protection. 2018;120:339-347. https://doi.org/10.1016/j.psep.2018.09.015
- [Google Scholar]
- Enhanced photocatalytic performance of hausmannite Mn3O4-rGO nanocomposite in degrading methylene blue. Materials Letters. 2021;305:130750. https://doi.org/10.1016/j.matlet.2021.130750
- [Google Scholar]
- Green synthesis of ag nanoparticles for plasmon-assisted photocatalytic degradation of methylene blue. Catalysts. 2021;11:1499. https://doi.org/10.3390/catal11121499
- [Google Scholar]
- Synthesis of multi-walled carbon nanotubes decorated with silver metallic nanoparticles as a catalytic degradable material via pulsed laser ablation in liquid media. Colloids and Surfaces A: Physicochemical and Engineering Aspects. 2021;626:126992. https://doi.org/10.1016/j.colsurfa.2021.126992
- [Google Scholar]
- Impact of irrigation with contaminated river water on growth, yield, and heavy metals accumulation in planted armenian cucumber (Cucumis melo var. flexuosus (L.) Naudin.) Water, Air, & Soil Pollution. 2025;236:1. https://doi.org/10.1007/s11270-024-07642-z
- [Google Scholar]
- The superior adsorption capacity of phenol from aqueous solution using Modified Date Palm Nanomaterials: A performance and kinetic study. Arabian Journal of Chemistry. 2022;15:104120. https://doi.org/10.1016/j.arabjc.2022.104120
- [Google Scholar]
- Surface modification of PP and PBT nonwoven membranes for enhanced efficiency in photocatalytic mb dye removal and antibacterial activity. Polymers. 2023;15:3378. https://doi.org/10.3390/polym15163378
- [Google Scholar]
- Poly (Vinyl Alcohol) hydrogels boosted with cross-linked chitosan and silver nanoparticles for efficient adsorption of congo red and crystal violet dyes. Gels (Basel, Switzerland). 2023;9:882. https://doi.org/10.3390/gels9110882
- [Google Scholar]
- Sustainable and green synthesis of carbon nanofibers from date palm residues and their adsorption efficiency for eosin dye. Sustainability. 2023;15:10451. https://doi.org/10.3390/su151310451
- [Google Scholar]
- Cyrene-enabled green electrospinning of nanofibrous graphene-based membranes for water desalination via membrane distillation. ACS Sustainable Chemistry & Engineering. 2024;12:17713-17725. https://doi.org/10.1021/acssuschemeng.4c06363
- [Google Scholar]
