Next Article in Journal
Post-Synthesis Ion Beam Sputtering of Pt/CeO2–ZrO2 Catalysts: Correlating Surface Modifications with Light-Off Performance
Previous Article in Journal
Facile Synthesis of Uniform NiO Nanoparticles Exclusively Confined in Mesoporous SBA-15 with High Loading for Ammonia Decomposition
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Sustainable Photocatalytic Treatment of Real Pharmaceutical Wastewater Using a Novel ZnO/MIP-202(Zr) Bio-MOF Hybrid Synthesized via a Green Approach

1
Environmental Engineering Department, Faculty of Engineering, Egypt-Japan University of Science and Technology (E-JUST), New Borg El-Arab City 21934, Egypt
2
Chemistry Department, College of Science, Imam Mohammad Ibn Saud Islamic University (IMSIU), Riyadh 11623, Saudi Arabia
3
Department of Chemistry, University College in Al-Qunfudhah 21912, Umm Al-Qura University, Makkah 24381, Saudi Arabia
4
Department of Chemistry, Faculty of Science, Islamic University of Madinah, Madinah 42351, Saudi Arabia
5
Public Works Engineering Department, Faculty of Engineering, Mansoura University, Mansoura 35516, Egypt
6
Chemical and Petrochemical Engineering Department, Faculty of Engineering, Egypt-Japan University of Science and Technology (E-JUST), New Borg El-Arab City 21934, Egypt
*
Authors to whom correspondence should be addressed.
Catalysts 2025, 15(11), 1017; https://doi.org/10.3390/catal15111017
Submission received: 18 September 2025 / Revised: 9 October 2025 / Accepted: 14 October 2025 / Published: 30 October 2025
(This article belongs to the Section Photocatalysis)

Abstract

Metal–organic frameworks (MOFs) are promising materials for environmental remediation, particularly in photocatalysis. In this work, a novel ZMIP nanocomposite was fabricated by integrating MIP-202(Zr) bio-MOF with ZnO nanoparticles. For the first time, ZnO nanoparticles were green-synthesized using water lettuce extract and incorporated into MIP-202(Zr) via a mild hydrothermal route. The resulting hybrid was applied as a visible-light photocatalyst for carbamazepine (CBZ) degradation in real pharmaceutical wastewater. Structural analyses (XRD, FTIR, TEM, EDS) verified the successful incorporation of ZnO into the MIP-202(Zr) framework. The composite exhibited a narrowed bandgap of 2.74 ± 0.1 eV compared to 4.05 ± 0.06 eV for pristine MIP-202 and 3.77 ± 0.04 eV for ZnO, highlighting enhanced visible-light utilization in ZMIP. Operational parameters were optimized using response surface methodology, where CBZ removal reached 99.37% with 84.39% TOC mineralization under the optimal conditions (90 min, pH 6, 15 mg/L CBZ, 1.25 g/L catalyst). The catalyst maintained stable performance over five reuse cycles. Radical quenching and UHPLC-MS analyses identified the dominant reactive oxygen species and generated intermediates, elucidating the degradation mechanism and pathways. Beyond CBZ, the ZMIP photocatalyst effectively degraded other pharmaceuticals, including doxorubicin, tetracycline, paracetamol, and ibuprofen, achieving degradation efficiencies of 82.93%, 76.84%, 72.08%, and 67.71%, respectively. Application on real pharmaceutical wastewater achieved 78.37% TOC removal under the optimum conditions. Furthermore, the supplementation of the photocatalytic system by inorganic oxidants ameliorated the degradation performance, following the order KIO4 > K2S2O8 > KHSO5 > H2O2. Overall, ZMIP demonstrates excellent activity, reusability, and versatility, underscoring its potential as a sustainable photocatalyst for real wastewater treatment.

Graphical Abstract

1. Introduction

The accelerated production and consumption of drugs for overcoming diseases in humans and animals have resulted in the detection of refractory pollutants in water streams [1]. Carbamazepine (CBZ), an antiepileptic and psychotropic drug, is commonly found in water systems due to its massive global consumption (1000 tons/year) [2]. Further, CBZ can reach water systems in different concentrations from ng/L to µg/L through wastewater treatment plants, livestock farming, and hospitals’ wastewater [3]. Hai et al. [4] mentioned that the CBZ concentration ranged from 30 to 6300 ng/L in the effluent of wastewater treatment plants in different countries such as Canada, Australia, and Finland. The accumulation of CBZ in water sources due to its bio-recalcitrance can harm aquatic organisms such as fish, algae, and invertebrates [5]. To mitigate the release of CBZ into the environment, various conventional biological, chemical, and physical treatment methods have been employed, including bacterial and fungal biodegradation, adsorption, electrocoagulation, and membrane filtration. However, their limited efficiency, high energy demand, high cost, and the generation of secondary pollution hinder their practical application [6]. Therefore, the development of an efficient, simple, cost-effective, and environmentally friendly degradation system is essential to mitigate the release of CBZ into aquatic environments.
Advanced oxidation processes (AOPs) such as Fenton, ozonation, and photocatalysis have been recently utilized for the degradation of bio-resistant pollutants (e.g., drugs) [7,8,9]. However, Fenton process necessitates acidic conditions and produces sludge, while ozonation process is expensive and generates toxic by-products. Regarding photocatalysis, it is a promising technique due to its simplicity, effectiveness, inexpensiveness, and green nature [10]. In photocatalysis, electron-hole (e/h+) pairs can be generated through the excitation of a semiconductor (e.g., TiO2, ZnO) by a light source with suitable energy leading to the production of reactive oxygen species (ROS) such as hydroxyl radicals (OH) and superoxide radicals (O2• −) and that can effectively degrade organic pollutants [11]. ZnO is featured by its high chemical stability, environmental friendliness, availability, low-cost, large excitation binding energy, and high electron mobility; therefore, it is a promising metal oxide for the employment in different applications such as CO2 reduction, water splitting, and photocatalytic degradation [12]. However, the preparation of ZnO requires the consumption of toxic chemicals which harms the environment and obstructs the scalable production. Thus, in this study, plant extracts were used as a reducing and stabilizing agents to replace dangerous chemicals which facilitates the practical production of ZnO. Water lettuce plant was used to fabricate ZnO due to its availability in water streams. The application of water lettuce in the green synthesis of ZnO can contribute to solving different environmental issues related to water lettuce such as blocking water streams, depletion of oxygen, increasing floods, and inhibiting biodiversity [13]. On the other hand, the large optical band gap (Eg) of ZnO and its inefficient charge carrier separation significantly limit its scalable application for the degradation of organic micropollutants, thereby hindering its photocatalytic performance [14]. These defects can be managed by different approaches such as metal doping, non-metal doping, fabricating heterojunctions, and decorating noble metals [15]. The fabrication of heterojunction can ameliorate the photocatalytic performance due to the charge transfer between the constituents of the heterostructure leading to the improvement of charge carrier separation. Oliveira et al. [16] synthesized a Bi2MoO6/ZnO heterojunction for the degradation of methylene blue dye indicating the improvement of the degradation efficiency to 99.7% in the case of the heterojunction compared to around 80% for pure ZnO. Nonetheless, the construction of heterojunction between ZnO and metal–organic frameworks (MOFs) has been rarely explored in the literature, and the degradation mechanism remains poorly investigated.
Recently, MOFs which are made of metal ions and organic ligands have been employed as photocatalysts due to their unique properties such as high surface area, porosity, and stability [17,18]. However, MOFs have some defects such as wide Eg, fast reunite of electron-hole pairs, and high cost due to the need for harmful, high-cost chemicals and high temperatures during the synthesis [19]. In this study, a zirconium-based bio-metal–organic framework (MIP-202(Zr) bio-MOF) was synthesized, where L-aspartic acid was employed as an organic ligand without using harmful chemicals and under low temperature [20]. However, the photocatalytic performance of MIP-202 is restricted due to its wide Eg and fast recombination rate. The integration of ZnO with MIP-202 can effectively suppress charge carrier recombination through interfacial transfer of electron–hole pairs between the two materials. Moreover, this interaction may lead to the formation of new energy levels during composite fabrication, thereby extending light absorption into the visible region. These synergistic effects enhance degradation kinetics and contribute to achieving higher mineralization efficiency [21]. Zhang et al. [22] fabricated a heterojunction of ZnO@NH2-MIL-88B showing the decrease in Eg to 2.22 eV in the case of ZnO@NH2-MIL-88B compared to 3.19 eV in the case of pure ZnO and the improvement of charge carrier separation which resulted in the enhancement of degradation performance for ZnO@NH2-MIL-88B compared to pure ZnO. In a recent study, C- and N-co-doped ZnO photocatalysts were synthesized via pyrolysis using ZIF-8 as the precursor. Optical characterization revealed that the band gap of pristine ZnO was approximately 3.37 eV, whereas the co-doped ZnO (C/N-ZnO-500) exhibited a reduced band gap, indicative of enhanced absorption within the visible-light region [23]. The coupling between ZnO and MIP-202 for the application in photocatalytic degradation of organic pollutants has not been explored in the previous studies, and the investigation of the degradation mechanism can provide new insights into the literature.
In this study, a novel ZnO/MIP-202(Zr) (ZMIP) heterostructure was fabricated by integrating bio-derived MIP-202(Zr) with ZnO nanoparticles through a mild hydrothermal process. To the best of our knowledge, this represents the first report on the construction of such a heterostructure. Furthermore, ZnO nanoparticles were synthesized via a green and sustainable route using water lettuce extract, providing an eco-friendly alternative to conventional chemical synthesis. This dual innovation highlights both the novelty in material design and the cost-effective, sustainable direction of the proposed photocatalyst. The synthesized ZMIP photocatalyst was thoroughly characterized using XRD, FTIR, TEM, and EDS to validate its structural integrity, morphological features, and elemental composition, thereby confirming the successful incorporation of ZnO into the MOF. In addition, the optical properties, including the band gap (Eg) and the relative positions of the conduction band (CB) and valence band (VB), were examined to assess the material’s light absorption capacity and charge transfer potential. The photocatalytic activity of ZMIP was then systematically investigated under visible light irradiation using CBZ as a model pharmaceutical contaminant. The influence of operational parameters such as irradiation time, pH, initial CBZ concentration, and catalyst dosage was optimized through response surface methodology (RSM), ensuring a statistically robust evaluation of the system. The kinetics of CBZ degradation were examined by applying pseudo-first-order and pseudo-second-order models, while total organic carbon (TOC) mineralization was assessed to evaluate the extent of organic load reduction as an additional indicator of the treatment performance. Catalyst durability and reusability were further studied through multiple repeated photocatalytic cycles. The underlying degradation mechanism was explored by combining optical property analysis with radical quenching experiments to clarify charge transfer pathways and identify the dominant reactive oxygen species, respectively. In addition, the degradation pathways were elucidated by detecting and characterizing transformation intermediates using an ultra-high-performance liquid chromatography coupled with mass spectrometry (UHPLC-MS) to provide insights into the sequential breakdown of CBZ. To explore its versatility and practical potential, the ZMIP/light system was further applied to the degradation of structurally diverse pharmaceuticals (doxorubicin hydrochloride, tetracycline, paracetamol, and ibuprofen) and tested across different water matrices (deionized water, tap water, lake water, seawater, and drain water) and real industrial wastewater collected from untreated pharmaceutical discharge. In addition, the influence of typical coexisting constituents, including natural organic matter (NOM, e.g., humic acid), inorganic anions (Cl, HCO3, NO3, SO42−), and inorganic cations (Mg2+, Ca2+, Na+, Cu2+), on CBZ degradation within the ZMIP/light system was systematically examined. Finally, the incorporation of highly reactive inorganic oxidants was examined to evaluate their effectiveness in mitigating electron–hole recombination and promoting the generation of reactive oxygen species. The tested inorganic oxidants were hydrogen peroxide (H2O2), potassium monopersulfate (KHSO5), potassium peroxydisulfate (K2S2O8), and potassium periodate (KIO4). Collectively, this work provides the first demonstration of a green-synthesized ZnO/MIP-202(Zr) heterostructure as a multifunctional, durable, and sustainable photocatalyst, supported by comprehensive experimental investigations ranging from synthesis and characterization to mechanistic studies and real-world applicability.

2. Results and Discussion

2.1. Physicochemical Characteristics of the Synthesized Materials

2.1.1. X-Ray Diffraction (XRD)

Figure 1a presents the XRD pattern of the green-synthesized ZnO nanoparticles, which shows excellent agreement with the reference data for the hexagonal wurtzite phase of zinc oxide (JCPDS No. 00-36-1451) [24]. Distinct diffraction signals were observed at 2θ angles of 31.69°, 34.36°, 36.17°, 47.53°, 56.52°, 62.80°, 66.30°, 67.97°, and 69.38°, corresponding to the (100), (002), (101), (102), (110), (103), (200), (112), and (201) crystallographic planes, respectively. These reflections confirm the successful synthesis of ZnO with a well-defined crystalline wurtzite structure. As shown in Figure 1b, the XRD pattern of the synthesized MIP-202(Zr) bio-MOF exhibits distinct diffraction peaks at 2θ values of 8.63°, 9.74°, 19.73°, and 21.59°, which can be indexed to the (111), (200), (420), and (440) crystallographic planes, respectively [25]. The sharp and well-defined reflections confirm the successful synthesis of the MIP-202(Zr) framework with a high degree of crystallinity. Additionally, the obtained XRD pattern of the synthesized MOF was compared with literature-reported data, as summarized in Table S1. The strong agreement between the observed and reported diffraction peaks verifies that the synthesized material possesses the characteristic crystal structure of MIP-202(Zr). On the other hand, the XRD pattern of the ZMIP photocatalyst (Figure 1c) displays characteristic diffraction peaks at 2θ values of 31.88°, 34.64°, 36.04°, 47.92°, 56.23°, 62.52°, 66.40°, 68.80°, and 69.70°, which are assigned to the (100), (002), (101), (102), (110), (103), (200), (112), and (201) planes of ZnO. Additional reflections at 8.12°, 9.74°, 13.27°, 20.14°, and 21.48° are attributed to the (111), (200), (222), (420), and (440) planes of MIP-202(Zr). The coexistence of diffraction features from both ZnO and MIP-202(Zr) demonstrates the successful formation of the ZMIP heterostructure, confirming the effective incorporation of the bio-MOF with the ZnO nanoparticles. Furthermore, for clarity and ease of comparison, all XRD patterns have also been compiled into a single combined figure (Figure S1).

2.1.2. Fourier-Transform Infrared (FTIR) Spectroscopy

Figure 2 presents the FTIR spectrum of the ZnO nanoparticles, MIP-202(Zr) bio-MOF, and ZMIP nanocomposite. In the FTIR spectrum of the synthesized ZnO nanoparticles, a characteristic absorption peak at 471.98 cm−1 is assigned to the Zn–O stretching vibration [26]. This peak confirms the successful formation of ZnO nanoparticles. In addition, the spectrum displays a broad and intense band at 3412.88 cm−1, which is attributed to the O–H stretching vibrations [27]. Further absorption features are observed at 2920.96 and 1408.69 cm−1, corresponding to C–H stretching and bending vibrations, respectively [28]. The band located at 1584.83 cm−1 is ascribed to aromatic C–C stretching [29], while the peak at 1049.58 cm−1 may be related either to C–N stretching of primary amines or to C–O stretching of primary alcohols [29].
The FTIR spectrum of the synthesized MIP-202(Zr) bio-MOF exhibits a broad absorption band centered at 3204.62 cm−1, which is attributed to the asymmetric and symmetric stretching vibrations of the −NH2 groups originating from L-aspartic acid [30]. The bands observed at 1610.3 and 1423.6 cm−1 correspond to the asymmetric and symmetric stretching modes of C−O bonds within the bio-organic linker, respectively [31]. A distinct absorption at 1234.77 cm−1 is attributed to C−N stretching vibrations [25]. Furthermore, the peaks at 647.61 and 482.27 cm−1 are assigned to Zr−COO and Zr−O bonds, respectively [31]. Collectively, these characteristic vibrational features provide compelling evidence for the successful assembly of the MIP-202(Zr) bio-MOF.
The FTIR spectrum of the ZMIP nanocomposite displays a broad absorption band at 3015.94 cm−1, which is attributed to the O–H stretching vibrations of the green-synthesized ZnO nanoparticles. This band partially overlaps with the −NH2 stretching modes arising from the MIP-202(Zr) bio-MOF, indicating successful integration of the two components within the composite structure. The band at 1600.47 cm−1 corresponds to the asymmetric C–O stretching vibration of the bio-organic linker, while the absorption at 1505 cm−1 is assigned to aromatic C–C stretching in ZnO. In addition, the signal observed at 1419 cm−1 is associated with C–H bending vibrations of ZnO, which overlaps with the symmetric C–O stretching of the MIP-202(Zr) linker. A distinct band at 1250.17 cm−1 is ascribed to C–N stretching vibrations within the MIP-202(Zr) framework, and the peak at 988.63 cm−1, slightly shifted, may correspond to either C–N stretching of primary amines or C–O stretching of primary alcohols in ZnO. Moreover, the absorption at 653.72 cm−1 is consistent with Zr–COO bonding in the MIP-202(Zr) MOF, whereas the band at 463.94 cm−1 arises from the combined contributions of Zn–O vibrations in ZnO and Zr–O bonds in MIP-202. The slight shifts in peak positions relative to the individual components are attributed to the incorporation of ZnO nanoparticles into the MOF, which modifies the local electronic environment and influences hydrogen-bonding interactions of amine groups [32]. The FTIR results confirm the coexistence of characteristic functional groups from both components, validating the successful synthesis and effective integration of the ZMIP heterostructure.

2.1.3. Structural Morphology and Elemental Profiling

The transmission electron microscopy (TEM) micrographs of the green-synthesized ZnO nanoparticles at different magnifications (Figure 3a,b) revealed a predominantly spherical morphology with uniform spatial distribution. The particles exhibited an average diameter of approximately 9.06 nm, as further confirmed by the particle size distribution shown in Figure S2. The high-resolution TEM (HRTEM) image (Figure 3c) further confirmed their crystalline nature, as evidenced by distinct lattice fringes corresponding to the interplanar spacings characteristic of the hexagonal wurtzite phase of ZnO. This observation verifies the successful synthesis of ZnO nanoparticles. Moreover, the energy-dispersive X-ray spectroscopy (EDS) spectrum (Figure 3d) indicated a surface composition of 81.61 wt% Zn and 18.39 wt% O, consistent with the theoretical stoichiometry of ZnO. Complementary elemental mapping images (Figure 3e) demonstrated a homogeneous and well-integrated distribution of Zn and O throughout the nanoparticle matrix, confirming both compositional uniformity and structural integrity.
The TEM images of the MIP-202(Zr) bio-MOF at various magnifications (Figure 4a,b) revealed that the particles possessed a primarily spherical geometry with nanoscale features. The average particle diameter was estimated to be approximately 70.94 nm, as further confirmed by the particle size distribution presented in Figure S3. While the XRD profile of MIP-202 (Figure 1b) confirms its overall crystallinity, the HRTEM image (Figure 4c) reveals the presence of locally disordered or amorphous regions. Such features are frequently observed in MOFs and are often attributed to partial amorphization caused by electron beam irradiation, surface or defect-related heterogeneities, and the intrinsic dynamic transition between crystalline and amorphous domains [33]. Moreover, the EDS spectrum (Figure 4d) and elemental mapping (Figure 4e) of the MIP-202(Zr) bio-MOF provide further confirmation of its successful synthesis and uniform elemental composition. As shown in the EDS spectrum, the framework is mainly composed of Zr (43.8 wt%), C (34.18 wt%), O (11.5 wt%), Cl (8.86 wt%), and N (1.65 wt%). The elevated levels of Zr, C, and O are consistent with the presence of Zr-cluster nodes and bio-organic linkers, which constitute the fundamental structural units of the MIP-200 MOF. The detected Cl is attributed to residual chloride ions retained within the porous framework [31], whereas the nitrogen content originates from the −NH2 groups of the aspartic acid linker. Moreover, the elemental mapping images of the bio-MOF illustrate the homogeneous spatial distribution of Zr, C, O, Cl, and N, with the merged map, confirming their uniform dispersion across the framework. These observations indicate that the MIP-202(Zr) bio-MOF possesses a well-organized structure in which the key elements are evenly integrated, ensuring structural integrity and functional performance.
The TEM micrographs of the ZMIP nanocomposite reveal particles with irregular morphology and a distinctly porous architecture, as shown in Figure 5a,b. The HRTEM image (Figure 5c) further highlights the pronounced interfacial interactions between the ZnO nanoparticles and the MIP-202(Zr) bio-MOF, thereby confirming their effective integration at the nanoscale. The EDS spectrum (Figure 5d) indicates that the nanocomposite consists of Zr (62.09 wt%), C (11.29 wt%), Cl (7.7 wt%), and N (2.16 wt%) derived from the MIP-202(Zr) framework, while the detection of Zn (2.69 wt%) validates the incorporation of the ZnO phase. These results provide compelling evidence of the successful coexistence and structural integration of ZnO and MIP-202(Zr) within the ZMIP nanocomposite. Furthermore, the corresponding elemental mapping images (Figure 5e) corroborate this observation, showing a uniform and well-distributed spatial dispersion of Zr, C, O, Cl, Zn, and N throughout the nanohybrid.

2.2. Optical Performance of the Synthesized Materials

As shown in Figure 6, the estimated Eg values of the green-synthesized ZnO nanoparticles, MIP-202(Zr) bio-MOF, and the ZMIP nanocomposite were estimated from three replicate measurements, yielding average values of 3.77 ± 0.04, 4.05 ± 0.06, and 2.74 ± 0.1 eV, respectively. The associated error bars represent the experimental uncertainty. The synthesized ZnO nanoparticles and MIP-202(Zr) bio-MOF both exhibited a white coloration, whereas the ZMIP nanocomposite displayed a slightly different off-white appearance. This observable color variation indicates the formation of the composite and suggests a subsequent alteration in its bandgap properties. The pronounced reduction in Eg upon integrating ZnO nanoparticles into the MIP-202(Zr) framework demonstrates the successful modification of its electronic structure, leading to improved visible-light absorption and, consequently, enhanced photocatalytic performance [34]. The decrease in bandgap energy observed for the ZMIP nanocomposite relative to its individual constituents (ZnO and MIP-202(Zr) bio-MOF) can be ascribed to intense interfacial coupling and synergistic electronic interactions between the semiconductor phase (ZnO) and the MOF. Upon incorporation of ZnO nanoparticles into the MIP-202(Zr) matrix, the formation of a heterojunction interface promotes partial overlap and alignment of the conduction and valence band edges of both materials, thereby narrowing the overall effective bandgap [21,35]. Moreover, electronic delocalization and charge redistribution occurring at the ZnO–MOF interface may generate intermediate electronic states or defect levels, which diminish the energy barrier for electron excitation. In addition, the organic linkers derived from L-aspartic acid can enhance π–π conjugation and facilitate localized electronic transitions, collectively contributing to the apparent reduction in bandgap energy and improved visible-light absorption [36]. Similar behavior has been observed in previous studies. For instance, Parsa et al. [37] demonstrated that depositing ZnO onto the MIL-100(Fe) MOF effectively reduced the bandgap from 2.82 eV for pristine MIL-100(Fe) to 2.18 eV for the composite. Likewise, Harisankar et al. [38] reported that the ZnO@MOF-5 composite showed a reduced band gap of 3.7 eV, compared with 3.8 eV for pristine MOF-5 and 3.44 eV for pure ZnO.
Additionally, bandgap estimation using Tauc plots can be significantly affected by the presence of long absorption tails at low energies, which render the extracted Eg values highly sensitive to the selected fitting region. These tails are often attributed to sub-bandgap absorption processes associated with structural defects, surface disorder, or dangling bonds that introduce localized electronic states within the bandgap. Such defect-related states can act as trap levels, lowering the effective excitation energy required for electron transitions [39,40]. Additionally, the observed behavior may also be linked to Urbach tails, which arise from lattice disorder and phonon interactions, leading to an exponential absorption edge [41]. In the case of ZnO–MOF composites, the generation of defect states at the ZnO/MIP-202 interface, along with possible electronic perturbations induced by the organic linkers, could further contribute to the extended absorption tail. While these features complicate the precise determination of Eg, they also indicate enhanced light-harvesting ability, as the defect-mediated states facilitate absorption in the visible region. Thus, the presence of absorption tails, although challenging for bandgap fitting, may play a beneficial role in improving photocatalytic activity by broadening the spectral response of the catalyst [42,43].
Based on the estimated electronegativities of the MIP-202(Zr) bio-MOF (C24H34N6O32Zr6) and ZnO nanoparticles, which were 6.5 and 5.74 eV, respectively, the conduction band potentials (ECB) were calculated as –0.03 eV for the MIP-202(Zr) MOF and –0.65 eV for ZnO. The corresponding valence band potentials (EVB) were determined to be 4.02 eV and 3.12 eV, respectively. These findings reveal the distinct electronic structures of the two materials, suggesting that their combination can promote interfacial charge transfer within the composite system.

2.3. Contribution of Multiple Processes

Figure 7 illustrates the degradation efficiency of CBZ across different systems at pH 7, catalyst dosage of 1 g/L, initial CBZ concentration of 20 mg/L, and a reaction time of 60 min. The adsorption efficiency toward CBZ was first assessed under dark conditions. The green-synthesized ZnO nanoparticles, MIP-202(Zr) bio-MOF, and the ZMIP nanocomposite exhibited adsorption efficiencies of 50.09, 24.48, and 56.82%, respectively, confirming the superior adsorption efficiency of the composite material toward CBZ, resulting from its active surface sites and well-developed porous structure (Figure 5a,b). Under visible light irradiation, direct photolysis of CBZ was negligible, with only 7.26% removal in the light-only system. In contrast, the ZnO/light and MIP-202(Zr)/light systems achieved 65.18 and 30.38% CBZ degradation, respectively, whereas the ZMIP/light system reached 76.59%. In the presence of visible light, photoexcited electrons and holes are generated, which subsequently participate in redox reactions to produce ROS such as OH and O2•− radicals [11]. These ROS serve as key agents in the degradation of CBZ molecules, acting synergistically with the adsorption capacity of the green-synthesized materials. The ZMIP nanocomposite exhibited markedly higher photocatalytic activity under light illumination compared to the individual ZnO nanoparticles and MIP-202(Zr) bio-MOF, demonstrating the synergistic effect of combining ZnO nanoparticles with the MIP-202(Zr) bio-MOF in enhancing CBZ removal. This improvement is primarily ascribed to its reduced bandgap energy (Figure 6), which enhances visible-light absorption and thereby promotes the generation of a greater quantity of reactive species [44].

2.4. Model Generation, Optimization, and Validation

The relationship between the independent variables and the CBZ degradation efficiency in the ZMIP/light system was mathematically modeled in the Minitab® 22 statistical software and is described by the polynomial response function presented in Equation (1).
Y (%) = 95.2 − 0.277 D + 9.39 F − 4.42 G − 7.2 H + 0.00322 D2 − 0.7596 F2 − 0.0316 G2 + 7.5 H2 + 0.0148 DF + 0.0171 DG − 0.286 DH − 0.0069 FG − 0.98 FH + 2.84 GH
where D represents the reaction time (min), F is the solution pH, G denotes the initial CBZ concentration (mg/L), and H indicates the photocatalyst dosage (g/L).
Table 1 provides the full design matrix, incorporating both actual and coded levels of the four independent variables considered in this study. It also lists the corresponding experimental CBZ removal efficiencies together with the values predicted by the quadratic response surface model. A total of 29 experimental runs were conducted under different photocatalytic conditions within the ZMIP/light system to establish the reliability and predictive accuracy of the model. The experimentally measured CBZ removal efficiencies exhibited considerable variation, ranging from 48.9 to 96.25%, thereby covering a broad operational window and ensuring a comprehensive evaluation of the system performance. The close correspondence between experimental and predicted values further highlights the adequacy of the developed model in describing the degradation process.
The optimal operating conditions were determined using the response optimizer in the Minitab® 22 statistical software based on maximizing the CBZ removal efficiency as a target. The resulting optimum values were an initial CBZ concentration of 15 mg/L, a photocatalyst dosage of 1.25 g/L, pH 6, and a reaction time of 90 min, as depicted in Table 2. To validate the reliability of the quadratic polynomial model, a confirmatory experiment was performed under these conditions, yielding a removal efficiency of 99.37% (Figure 7). This closely matched the predicted value of 98.84%, with a minimal deviation of 0.53%. The strong agreement between experimental data and predicted values demonstrates the robustness of the quadratic polynomial model in predicting CBZ photodegradation within the ZMIP/light system and its effectiveness in optimizing operational variables.
In addition, the degradation efficiency of CBZ in the ZMIP/light system was evaluated against the performance of a range of photocatalytic systems incorporating different metal oxide–MOF composite photocatalysts for the elimination of various pharmaceutical organic contaminants, as presented in Table 3. The results highlight the superior efficacy of the proposed system.

2.5. Analysis of Variance

ANOVA was carried out using the Minitab® 22 statistical software to evaluate the adequacy and reliability of the developed response surface model, as explained in Table 4. Being very close to unity, the R2 (97.88%) and adjusted R2 (95.76%) highlight the robustness and reliability of the regression, confirming that the model can accurately describe the experimental data. The close agreement between the two coefficients, with only a 2.12% discrepancy, indicates strong consistency between the predicted and experimental responses, thereby confirming the statistical validation of the fitted polynomial model in estimating the influence of experimental variables on CBZ removal efficiency in the ZMIP/light system. In addition, the high R2 value indicates that 97.88% of the variation in CBZ removal efficiency can be explained by the model, whereas just 2.12% of the variation cannot be accounted for in the developed equation. The statistical adequacy and significance of the model was further verified by the high F-value (46.18) and the very low p-value (< 0.05). This is supported by the minimal difference observed between the predicted and experimental responses (Table 2). Moreover, several model terms (D, F, G, H, DF, DG, DH, FG, FH, GH, D2, F2, G2 and H2) were identified as statistically significant, as their p-values were less than 0.05.

2.6. Influence of Operational Factors

The F-value was employed to assess the relative influence of each variable on the response. As presented in the ANOVA results (Table 4), the photocatalyst dose exhibited the highest F-value (167.33), considerably exceeding those of the other parameters. Accordingly, the magnitude of the variables’ effects on CBZ removal efficiency within the ZMIP/light system can be ranked as follows, photocatalyst dose > initial CBZ concentration > reaction time > solution pH, highlighting the dominant role of photocatalyst dosage in governing the degradation process.
To provide a clear visualization of how the interaction among the independent operational parameters (photocatalyst dosage, initial CBZ concentration, solution pH, and reaction time) affects CBZ degradation in the ZMIP/light system, contour plots were generated using the Minitab® 22 statistical software, with two variables represented as coordinate axes, as shown in Figure 8.
The effect of ZMIP dosage on CBZ degradation efficiency was evaluated within the range of 0.75–1.25 g/L, as explained in Table 1 and Table 5. The combined effects of reaction time and photocatalyst dosage on CBZ removal, under fixed conditions of pH 7 and an initial CBZ concentration of 20 mg/L, are illustrated in Figure 8a. The results demonstrated that CBZ removal efficiency increased proportionally with both photocatalyst dosage and reaction time. The highest degradation efficiency (96.25%) was achieved at a catalyst dosage of 1.25 g/L after 90 min. The improved CBZ degradation at the optimum dosage is attributed to the increased number of active sites on the photocatalyst surface, which enhances photon absorption and promotes higher ROS generation, thereby accelerating pollutant removal [56]. However, exceeding the optimal photocatalyst loading resulted in reduced degradation efficiency. This decline is primarily attributed to light scattering and turbidity at higher catalyst concentrations, which hinder effective light penetration and limit photon access to active sites, thereby suppressing charge carrier excitation [57]. In addition, particle agglomeration at higher dosages can further reduce the effective surface area and overall catalytic performance [58].
The influence of initial CBZ concentration on its degradation efficiency was investigated in the range of 15–25 mg/L at pH 7 and a ZMIP dosage of 1 g/L (Table 1 and Table 5). The contour plot showing the combined effects of CBZ concentration and reaction time is presented in Figure 8b. The results demonstrated a decline in removal efficiency with increasing initial CBZ concentration. The maximum degradation efficiency (93.12%) was achieved at 15 mg/L after 90 min. At low CBZ concentrations, the generated ROS are sufficient to degrade pollutants efficiently. However, as CBZ concentration increases, the available ROS become inadequate to sustain complete degradation, thereby reducing efficiency and requiring longer reaction times to reach comparable performance [59]. In contrast, since the number of active sites on the catalyst surface remains fixed, excess CBZ molecules (above the optimum value of 15 mg/L) can saturate these sites rapidly and suppress ROS generation, thereby hindering the degradation process [60]. In addition, at higher CBZ concentrations, pollutant molecules can absorb incident photons, thereby reducing the light available for electron–hole pair generation and reducing the CBZ removal efficiency [61].
The impact of solution pH on CBZ degradation efficiency was investigated in the range of 3–11 at a ZMIP dosage of 1 g/L and an initial CBZ concentration of 20 mg/L (Table 1 and Table 5). The combined effects of pH and reaction time on CBZ removal are presented in Figure 8c. Since the pKa of CBZ is approximately 13.9, it remains essentially neutral across the tested pH range (3–11) [62]. Therefore, its adsorption onto the ZMIP surface is not driven by electrostatic interactions but rather by π–π stacking, hydrophobic interactions, and hydrogen bonding [63]. On the other hand, the pHpzc of the ZMIP photocatalyst was determined to be 8.6 (Figure S4), indicating a positively charged surface at pH < 8.6 and a negatively charged surface at pH > 8.6 [64]. The photocatalytic performance followed the order pH 3 > pH 7 > pH 11, with the highest degradation efficiency (79.32%) obtained at pH 3 after 90 min. At pH 3, the ZMIP surface is positively charged while CBZ remains neutral, enabling interactions through hydrogen bonding (via the amide group) and π–π stacking [65]. In addition, the positively charged surface enhances hydroxyl ion (OH) adsorption near the active sites of the photocatalyst, where they are oxidized by photogenerated holes to form OH radicals, thus improving degradation efficiency under acidic conditions [66]. At neutral pH 7, the catalyst surface remains positively charged, but less strongly than at pH 3. CBZ also remains neutral, allowing adsorption through π–π stacking and hydrogen bonding, though to a lesser extent than under acidic conditions. As a result of the lower proton concentration, OH generation at pH 7 is only moderate, leading to a lower CBZ degradation efficiency compared with pH 3. In contrast, at pH 11, the catalyst surface becomes strongly negatively charged (pH > pHpzc), while CBZ remains neutral. The negative surface charge repels OH ions from the catalyst surface, thereby reducing their availability for oxidation by photogenerated h+ to produce OH radicals. In addition, the negative surface potential hinders electron transfer from the conduction band to electron acceptors in solution, leading to electron accumulation within the catalyst or at surface states. This accumulation promotes electron–hole recombination and suppresses ROS generation [67]. Additionally, weaker hydrogen bonding and diminished surface–substrate interactions under alkaline conditions further reduce photocatalytic efficiency [68].
As shown in Figure 8, extending the irradiation time from 30 to 90 min significantly enhanced CBZ degradation efficiency. This enhancement is due to the prolonged exposure of photocatalyst active sites to light, which allows greater photon absorption. The longer irradiation also increases the generation of reactive species and photogenerated carriers. As a result, the probability of interactions between ROS and CBZ molecules becomes higher [69]. However, this positive effect persists only up to a certain duration and extended irradiation beyond the optimum time can decrease the overall degradation efficiency. Prolonged illumination periods may promote the recombination of photogenerated electron–hole pairs and reduce the delocalization of charge carriers, leading to an equilibrium between radical generation and recombination [70].

2.7. Photocatalytic Degradation Kinetics of CBZ

The degradation kinetics of CBZ in the ZMIP/light system were analyzed using the pseudo-first-order and pseudo-second-order kinetic models. The experiments were conducted at neutral pH (7), with a catalyst dosage of 1 g/L, and a reaction period of 60 min. Three initial CBZ concentrations, 15, 20, and 25 mg/L, were examined under these experimental conditions. As shown in Figure 9a, the apparent first-order rate constants (K1) for CBZ degradation were 0.0327 (R2 = 0.9936) min−1, 0.0254 (R2 = 0.9935), and 0.022 (R2 = 0.9828) min−1 at initial CBZ concentrations of 15, 20, and 25 mg/L, respectively. In addition, the apparent second-order rate constants (K2) were 0.0069 (R2 = 0.9792) min−1, 0.0036 (R2 = 0.9645), and 0.0025 (R2 = 0.9365) min−1, respectively, at the same CBZ concentrations, as presented in Figure 9b. The higher R2 obtained from the first-order model relative to the second-order model demonstrates that the photodegradation of CBZ in the ZMIP/light system is best fitted by first-order kinetics. The pseudo-first-order kinetic model is widely regarded as the most suitable framework for describing the photocatalytic degradation behavior of organic pollutants in aqueous environments [71]. Moreover, the results confirmed a decline in the CBZ degradation rate with increasing initial CBZ concentration. These findings are consistent with the previously discussed results regarding the effect of initial pollutant concentration on the system performance. On the other hand, the slight deviations from linearity observed in the kinetic plots likely arise from complex surface adsorption–desorption equilibria and the simultaneous occurrence of photogenerated electron–hole recombination during the photocatalytic reaction. These processes influence the apparent degradation rate and are not fully captured by simplified kinetic models that assume uniform active site distribution and ideal reaction conditions. Nevertheless, the pseudo-first-order kinetic model remains a widely accepted and effective empirical approach for describing and comparing the photocatalytic degradation behavior of organic contaminants in heterogeneous systems [72,73].

2.8. The Reusability of the ZMIP Photocatalyst

The reusability of the ZMIP photocatalyst was assessed under the optimum conditions (pH 6, catalyst dosage of 1.25 g/L, initial CBZ concentration of 15 mg/L, and reaction duration of 90 min) by recovering and reutilizing the ZMIP composite powder after each cycle over five consecutive photocatalytic runs. Following each photodegradation cycle, the ZMIP powder was collected and subjected to repeated centrifugation (EBA 20, Hettich, Beverly, MA, USA) with ultrapure water and ethanol until the supernatant was entirely transparent. The recovered material was then dried at 60 °C for 24 h in an electric oven (ED 53, Binder, Tuttlingen, Germany) to eliminate residual moisture. The dried photocatalyst was subsequently introduced into a newly prepared TCN solution to commence the next experimental run [74,75]. As illustrated in Figure 10a, the nanocomposite demonstrated only a slight decline in CBZ removal efficiency, achieving 99.37, 96.80, and 94.54% after the first, second, and third successive cycles, respectively. At the end of the fourth and fifth cycles, the degradation efficiency decreased to 85.23 and 77.97%, respectively, indicating that prolonged reaction times might be necessary in the later stages to achieve the desired pollutant degradation. The gradual decline in the catalytic performance of the ZMIP photocatalyst over successive cycles can be ascribed to multiple factors, including partial catalyst loss during recovery, diminished surface area due to potential nanoparticle agglomeration, leaching of Zn2+ ions into the reaction medium, and progressive surface fouling caused by the adsorption of CBZ molecules and their degradation intermediates onto the active sites [76,77]. Despite these challenges, the ZMIP composite retained sufficient photocatalytic activity, confirming sustained reactive species generation under continuous light exposure and highlighting its potential for long-term and practical environmental applications.

2.9. Total Organic Carbon (TOC) Mineralization in the ZMIP/Light System

The efficiency of TOC removal during the photocatalytic degradation of CBZ using the ZMIP photocatalyst was evaluated under the optimum conditions as an additional indicator of the treatment performance. At an initial CBZ concentration of 15 mg/L and an initial TOC concentration of 6.24 mg/L, the system achieved a TOC mineralization efficiency of 84.39% after 90 min of irradiation, as presented in Figure 10b. These results confirm the promising potential of the ZMIP photocatalyst for application in real industrial wastewater treatment. Although TOC removal progressively increased over time during the photocatalytic reaction, the overall TOC mineralization efficiency was lower than the CBZ removal efficiency under identical conditions (Figure 7), owing to the sequential formation of organic intermediates derived from CBZ [78]. The presence of these organic intermediates extends the persistence of organic load in the reaction medium and competes with the parent pollutant for the generated ROS, thereby incomplete TOC mineralization was attained [79]. Complete mineralization of CBZ to environmentally benign end products (CO2 and H2O) requires prolonged reaction times to enable repeated attacks by ROS on the intermediate degradation products [80].

2.10. Photocatalytic Degradation Mechanism of CBZ

Considering the calculated ECB and VCB potentials of the synthesized ZnO nanoparticles and MIP-202(Zr) bio-MOF, a plausible charge transfer pathway within the ZMIP/light photocatalytic system is proposed. Upon illumination, photons excite the ZMIP composite, promoting electrons from the VB to the conduction band CB and generating corresponding holes in the VB. This process creates e/h+ pairs within both components of the composite, initiating charge separation [81]. Subsequently, the photogenerated electrons in the CB of ZnO migrate to the CB of the MIP-202(Zr) MOF, while the holes in the VB of MIP-202(Zr) are transferred to the VB of ZnO, facilitating effective charge separation across the ZMIP heterojunction [79]. The interfacial charge transfer between the green-synthesized ZnO nanoparticles and MIP-202(Zr) bio-MOF effectively suppresses the recombination of photogenerated e/h+ pairs, thereby prolonging the lifetime of charge carriers and enhancing the photocatalytic degradation efficiency of the target pollutant [82]. Moreover, the photogenerated electrons can react with dissolved molecular oxygen to form O2• − radicals, while the holes in the VB could oxidize H2O molecules and/or OH ions to generate hydroxyl radicals (OH) [83]. Subsequently, the generated O2• − radicals can react with protons (H+) to produce H2O2, which may further interact with electrons to yield additional OH radicals [84]. The photogenerated holes, O2• − radicals, and OH radicals can interact with organic molecules adsorbed near the active sites on the ZMIP photocatalyst surface, thereby enhancing the oxidative degradation of the targeted organic contaminant [85]. Figure 11 provides a schematic representation of the charge transfer pathway between MIP-202(Zr) framework and ZnO nanoparticles, along with the proposed mechanism for ROS generation resulting from their interaction.
To quantify the participation of the generated ROS in the photocatalytic reaction within the ZMIP/light system, ammonium oxalate, benzoquinone, and tert-butyl alcohol were employed to deactivate the oxidation ability of h+, O2• − radicals, and OH radicals, respectively [86,87]. The experiments were employed under the optimum conditions (pH 6, catalyst dosage of 1.25 g/L, initial CBZ concentration of 15 mg/L, and reaction duration of 90 min). As demonstrated in Figure 12a, the degradation efficiency of CBZ was reduced to 82.9, 68.65, and 50.15% upon the addition of 1 mM of ammonium oxalate, benzoquinone, and tert-butyl alcohol, respectively, compared to 99.37% without any quenching agents. These results demonstrated that all types of ROS contributed to CBZ photocatalytic decomposition process, confirming the oxidation mechanism. However, the dominant reactive species in CBZ degradation within the ZMIP/light system followed the order OH > O2• − > h+, indicating that OH radicals were the most effective contributors, while photogenerated holes played the least significant role.

2.11. Degradation of Other Organic Pollutants

A systematic evaluation of the photocatalytic performance of the ZMIP/light system was carried out for the degradation of various pharmaceutical contaminants, including doxorubicin hydrochloride, tetracycline, paracetamol, and ibuprofen. Each compound was investigated separately under the operating conditions optimized for CBZ degradation (pollutant concentration = 15 mg/L, pH = 6, ZMIP dosage = 1 g/L, and reaction time = 90 min). The degradation efficiencies were 82.93% for doxorubicin hydrochloride, 76.84% for tetracycline, 72.08% for paracetamol, and 67.71% for ibuprofen, in comparison with 99.37% observed for CBZ, as presented in Figure 12b. These outcomes highlight the capability of the ZMIP photocatalyst to effectively degrade a broad spectrum of emerging pharmaceutical pollutants, demonstrating not only the potential of the ZMIP/light system in practical water purification applications but also the versatility of the operating conditions originally optimized for CBZ. The observed variations in degradation efficiencies among different compounds can be primarily attributed to the fact that the applied operating parameters were optimized for CBZ removal, emphasizing the necessity of customizing catalytic conditions to maximize pollutant-specific degradation. Furthermore, detailed investigation into the factors governing pollutant selectivity and the intrinsic limitations of ZMIP would provide valuable insights for improving the catalyst’s applicability in diverse water treatment contexts.

2.12. CBZ Proposed Degradation Paths

Based on the detected m/z peaks of CBZ intermediates formed during its degradation by the ZMIP photocatalyst under light irradiation (Figure S5), the corresponding by-products were identified, and the photocatalytic degradation pathways of CBZ were subsequently proposed, as illustrated in Figure 13. In Path A, the olefinic double bond within the heterocyclic ring of CBZ (m/z = 236) was initially attacked by superoxide and hydroxyl radicals, leading to its oxidation into the intermediate product P1 (m/z = 301) which then underwent intramolecular cyclization and dehydration to form the by-product P2 (m/z = 267) [88]. Through decarboxylation processes, the intermediate P3 (m/z = 224) was transformed into P4 (m/z = 196) [89]. Then, through successive oxidation steps involving photogenerated holes, iodate, hydroxyl, and superoxide radicals, the intermediate P4 was further transformed into the smaller aromatic P5 (m/z = 138) [90]. In Path B, CBZ underwent oxidative transformation initiated by superoxide radicals and the photogenerated holes, forming the by-product P6 (m/z = 253) [6]. The intermediate P6 was converted into P3 through ring contraction and cleavage reactions [91]. Similar to Path A, the by-product P3 may undergo further oxidation to produce the smaller degradation products P4 and P5. Alternatively, P3 can experience N–C bond cleavage induced by superoxide radicals, leading to the formation of P7 (m/z = 180), which can subsequently be oxidized to generate P4 and P5 [92]. In both proposed pathways, P5 was eventually mineralized into CO2 and H2O, highlighting the efficiency of the ZMIP/light system in facilitating the stepwise degradation of CBZ [46].
The MS data revealed that the peak at m/z = 267 (P2) exhibited the highest signal intensity among the detected intermediates, suggesting that the initial oxidation of the olefinic double bond, followed by intramolecular cyclization and dehydration, represents a major and kinetically favored transformation route. In contrast, the peak at m/z = 301 (P1), attributed to a hydroxylated intermediate, appeared with significantly lower intensity, indicating that this species is transient and rapidly converted to P2. The prominence of P2 (m/z = 267) following the parent compound CBZ (m/z = 236) confirms its substantial accumulation during degradation. Meanwhile, intermediates associated with Path B (e.g., P6, m/z = 253) were detected with noticeably lower relative intensities, implying that this route proceeds only to a limited extent. Based on the relative abundance and persistence of the intermediates, particularly those corresponding to m/z = 267, it can be concluded that Path A is the predominant degradation route under the experimental conditions. This pathway ultimately leads to the formation of smaller aromatic intermediates, which are further mineralized into CO2 and H2O.

2.13. Effect of Coexisting Substances

Real wastewater is characterized by a complex composition of inorganic ions and organic compounds, which may hinder the photocatalytic degradation of organic contaminants [93]. To better simulate practical application scenarios, various real water samples collected from different natural sources were utilized as reaction solutions for the degradation of CBZ within the ZMIP/light system. The evaluated matrices included deionized water, tap water, seawater, lake water, and drain water. The collection sites of the raw wastewater samples are presented in Table S2, whereas their corresponding physicochemical characteristics are summarized in Table S3. In addition, a detailed description of pre-processing and conditioning procedures of these samples is provided in Text S1. All experiments were performed under the optimum conditions: pH 6, catalyst dosage of 1.25 g/L, and initial CBZ concentration of 15 mg/L. As illustrated in Figure 14a, the deionized water yielded the highest CBZ degradation efficiency among all tested water matrices, achieving 99.37% after 90 min. Tap water also demonstrated a relatively high removal efficiency of 92.81%, showing only a 6.6% reduction compared to deionized water. In contrast, significantly lower degradation efficiencies were observed in seawater (70.86%), lake water (54.06%), and drain water (31.08%). The higher CBZ removal efficiencies observed in deionized and tap water matrices can be attributed to their relatively low matrix complexity, as evidenced by their low turbidity and minimal concentrations of dissolved organic matter and total dissolved inorganic ions [94]. In contrast, the reduced photocatalytic performance of the ZMIP photocatalyst in sea, lake, and drain water is likely due to the increased complexity of these matrices, which may interfere with light penetration and active site availability, thereby inhibiting the overall degradation efficiency [95,96]. Natural organic matter (NOM) can adversely affect photocatalytic degradation by competing with the target pollutant for ROS, thereby diminishing its removal efficiency [97]. Additionally, the high aromaticity of NOM, reflected by elevated specific ultraviolet absorbance (SUVA) values (≥2 L/mg.m), promotes significant light absorption. This competitive absorption diminishes the photons available to excite the photocatalyst and the target pollutant, thereby limiting irradiation at the active sites and ultimately reducing ROS generation [98]. Moreover, NOM may adsorb onto the surface of the photocatalyst, obstructing active sites and further hindering the production of ROS essential for pollutant degradation. On the other hand, the elevated turbidity in lake, drain, and seawater matrices may scatter incident light, thereby limiting the penetration of light to the catalyst’s active sites [99]. In addition, the presence of inorganic salts in these matrices can inhibit the photocatalytic process by acting as scavengers of ROS, leading to the formation of less reactive radicals and ultimately reducing the degradation efficiency of CBZ [100].
Additionally, the effects of common coexisting substances such as NOM (e.g., humic acid), inorganic anions (Cl, HCO3, NO3, SO42−), and inorganic cations (Mg+2, Ca+2, Na+, Cu+2) on the removal efficiency of CBZ in the ZMIP/light system were investigated. The initial CBZ concentration was set at 15 mg/L, with the solution pH maintained at 6, while the photocatalyst was introduced at a dosage of 1.25 g/L. After 60 min of reaction, the presence of humic acid at a concentration of 5 mg/L slightly enhanced the photocatalytic degradation of CBZ, increasing the removal efficiency to 81.5% compared to 76.59% in the absence of humic acid, as illustrated in Figure 14b. This suggests that the presence of a low level of humic acid can positively influence the photocatalytic process. Indeed, humic acid is a naturally occurring photosensitizer capable of absorbing light energy, thereby facilitating photochemical energy conversion and potentially promoting the generation of reactive species during the reaction [101]. However, a progressive inhibitory effect on CBZ degradation was observed with increasing concentrations of humic acid. Specifically, the removal efficiency declined to 67.84 and 59.23% at humic acid concentrations of 10 and 20 mg/L, respectively. This reduction is likely attributed to competitive interactions between humic acid and CBZ for ROS, which hampers the availability of ROS for CBZ oxidation [102]. These findings highlight the importance of removing excess humic substances, particularly at concentrations exceeding 5 mg/L, through a pre-treatment step when applying the proposed photocatalytic system for real wastewater treatment applications.
As shown in Figure 15a, the addition of 10 mM of Cl, HCO3, NO3, and SO42− resulted in a decrease in CBZ removal efficiency to 93.02, 85.76, 80.92, and 75.34%, respectively, after 90 min, compared to 99.37% in the absence of added anions. The results indicated that the Cl ions exhibited a negligible impact, while both HCO3 and NO3 significantly suppressed the photocatalytic activity. The most pronounced inhibition was observed with SO42−, indicating that the presence of sulfate in practical applications may substantially reduce the system’s degradation efficiency. Overall, the inhibitory effects of the tested anions on CBZ degradation within the ZMIP/light system followed the order: SO42− > NO3 > HCO3 > Cl. On the other hand, the addition of 2 mM of Mg+2, Ca+2, Na+, and Cu+2 reduced the CBZ removal efficiency to 96.02, 87.59, 82.39, and 70.30%, respectively, after 90 min of reaction, as presented in Figure 15b. The addition of Mg+2 had a negligible effect on CBZ removal, whereas the presence of Ca+2 and Na+ led to a significant decline in the photocatalytic efficiency of the ZMIP/light system. The Cu2+ ions exhibited the strongest inhibitory effect. These findings suggest that the photocatalytic performance of the proposed system may be significantly reduced in real wastewater containing Ca+2, Na+, and particularly Cu2+, while Mg+2 is unlikely to cause substantial interference. The decline in degradation efficiency in the presence of inorganic ions can be explained by their dual interaction with ROS. On one hand, these ions may react with ROS to generate secondary radicals, which generally exhibit lower oxidation potential and reactivity [103]. On the other hand, they can also act as quenchers that directly consume ROS, thereby reducing their overall availability. Both pathways ultimately diminish the oxidative capacity of the system [104]. Moreover, competition between inorganic ions and CBZ molecules for the limited reactive sites on the photocatalyst surface may further suppress the photocatalytic degradation performance [105].

2.14. Remediation of Real Industrial Discharge

Real pharmaceutical wastewater was treated using the ZMPI photocatalyst in the presence of visible light to assess the system’s practical photocatalytic performance. Due to the expected complex composition of the water sample, the reduction in TOC was selected as a representative indicator of overall treatment efficiency. The photocatalytic process was conducted over five successive cycles, each with a reaction time of 90 min, at pH 6 and ZMIP dosage of 1.25 g/L. As illustrated in Figure 16a, the TOC mineralization efficiencies of the real sample (initial TOC = 75.34 mg/L) declined from 78.37% in the first cycle to 72.24, 63.76, 50.50, and 38% after the repeated cycles, respectively. The results demonstrated a substantial reduction in TOC values, indicating effective degradation of the organic content. However, minimal TOC removal was observed during the initial stages of the reaction, likely due to the formation of intermediate by-products that retain organic carbon and compete with the parent organic compounds for ROS [106]. In the later stages of the reaction, further degradation of organic intermediates occurred, leading to a notable decrease in TOC levels [107]. However, complete mineralization requires the generation of sufficient ROS to fully degrade both the original organic pollutants and the intermediate by-products [97]. These findings underscore the promising potential of the ZMPI composite for the effective treatment of industrial wastewater.

2.15. Effect of Introducing Highly Reactive Inorganic Oxidizing Agents

The recombination of photogenerated e/h+ pairs is a major drawback in photocatalytic systems, as it results in significant energy loss and decreases the availability of charge carriers required for redox reactions [108]. To overcome this limitation, reactive inorganic oxidants such as H2O2, KHSO5, K2S2O8, and KIO4 were introduced into the reaction solution before initiating the photocatalytic reaction. These oxidants function as efficient electron acceptors by capturing photogenerated electrons on the photocatalyst surface. This process suppresses the recombination of e/h+ pairs, preserving more charge carriers for redox reactions. In addition, the presence of these oxidants facilitates the production of more ROS, which play a decisive role in accelerating the photocatalytic degradation process [109]. The experiments were carried out at pH 7, using a ZMIP dosage of 1 g/L and an initial CBZ concentration of 20 mg/L. Each oxidant was introduced separately into the reaction solution at a concentration of 3 mM. Following the addition of these components, the light lamp was switched on to initiate the photocatalytic reaction. The results, shown in Figure 16b, demonstrate that the addition of KIO4 produced the most pronounced improvement in CBZ degradation efficiency (98.93%), followed by K2S2O8 (92.15%), KHSO5 (89.22%), and H2O2 (81.86%). In contrast, the system without any added oxidant achieved a lower degradation efficiency of 76.59%.
The superior CBZ removal efficiency observed with KIO4, relative to other oxidants, is likely due to the comparatively extended I–O bond length in periodate ions (IO4) (approximately 1.78 Å), which enhances its susceptibility to activation and subsequently promotes the formation of various ROS [110]. Periodate ions possess a strong electron-accepting capability, enabling them to effectively capture photogenerated electrons on the surface of the ZMIP photocatalyst, as illustrated in Equation (2) [111]. This electron-scavenging process suppresses e/h+ recombination, thereby extending the lifetime of photogenerated holes and enhancing their availability for subsequent oxidative reactions. Moreover, upon visible-light illumination, IO4 ions can produce iodate radicals (IO3), as illustrated in Equation (3) [112]. In addition, IO4 plays a key role in the formation of O2• − radicals and singlet oxygen (1O2), as described in Equations (4) and (5) [113,114]. On the other hand, the oxygen-containing functional groups on the ZMIP nanocomposite surface, as indicated in the FTIR spectrum (Figure 2), can interact with IO4 ions to produce additional IO3 radicals as well as OH radicals, as demonstrated in Equations (6) and (7) [74]. These ROS actively participate in the oxidative degradation of CBZ, breaking it down into smaller intermediate compounds, CO2, and H2O [34].
The addition of K2S2O8 significantly enhanced the degradation efficiency of CBZ, primarily due to the ability of persulfate ions (S2O82−) to scavenge photogenerated electrons from the illuminated ZMIP photocatalyst, thereby suppressing e/h+ recombination and facilitating the formation of sulfate radicals (SO4• −) as potent ROS, as illustrated in Equation (8) [115]. Furthermore, the homolytic cleavage of the peroxide bond in S2O82− under light irradiation leads to the generation of additional SO4• − (Equation (9)) [116]. Hydroxyl radicals can also be produced via the reaction of SO4• − with H2O molecules or OH ions, as shown in Equations (10) and (11) [117,118]. In addition, other ROS, such as O2• − radicals and 1O2, may be formed through subsequent secondary reactions, as presented in Equations (12)–(14) [119,120]. Moreover, the oxygen-containing functional groups present on the surface of the photocatalyst can react with S2O82− ions, leading to the generation of additional SO4• − radicals, as illustrated in Equations (15) and (16) [75]. The slightly lower CBZ degradation efficiency with K2S2O8 compared to KIO4 may result from the quenching effect of S2O82− ions on SO4• − and OH radicals. This interaction converts the highly reactive radicals into the less active species (S2O8• −), as illustrated in Equations (17) and (18) [77].
Analogous to K2S2O8, the incorporation of KHSO5 into the ZMIP/light photocatalytic system enhanced CBZ degradation efficiency, primarily due to the dual role of peroxymonosulfate anion (HSO5) in accepting photogenerated electrons and concurrently producing SO4• −, as illustrated in Equation (19) [121]. In addition, HSO5 can undergo direct photolysis under light irradiation, resulting in the formation of both SO4• − and OH radicals, as shown in Equation (20) [122]. Further, O2• − radicals can be generated through the reaction between HSO5 ions and SO4• − radicals, as presented in Equations (21) and (22) [123,124]. Like the K2S2O8-based system, OH radicals and the 1O2 species can be generated through the reactions presented in Equations (10) and (11) and (Equation (14)), respectively. The functional groups on the surface of the ZMIP catalyst can also activate HSO5 to produce SO4• − radicals, as explained in Equations (23) and (24) [125,126].
Nevertheless, the CBZ removal efficiency achieved with KHSO5 was lower than that obtained with K2S2O8. This difference can be explained by the structural characteristics of the oxidants. The O–O bond length in the HSO5 ions (1.453 Å) is shorter than that in S2O82− ions (1.497 Å). As a result, S2O82− ions are more easily activated and more efficient in producing ROS [96].
Although H2O2 possesses the capability to scavenge photogenerated electrons and can be activated under visible light to produce OH radicals (Equations (25) and (26)) [122,123], its addition resulted in the lowest CBZ removal efficiency among the tested inorganic oxidants. This limited performance can be attributed to the relatively strong O–O bond in H2O2 molecules, which impedes its cleavage and activation by the photocatalyst, thereby restricting the generation of ROS [127]. Moreover, the lower oxidation potential and shorter half-life of OH radicals, dominant in the H2O2-based system, further constrain its oxidative effectiveness when compared to SO4• − radicals, which are efficiently generated in the K2S2O8 and KHSO5 systems [75].
I O 4 + 8 e + 8 H + H 2 O + I
I O 4 + h v O + I O 3
3 I O 4 + 2 O H H 2 O + 3 I O 3 + 2 O 2
I O 4 + 2 O 2 + H 2 O I O 3 + 2 O H + 2 ( O 2 1 )
ZMIP OOH + I O 4   ZMIP OO + OH + I O 3
ZMIP OH + I O 4   ZMIP O + OH + I O 3
S 2 O 8 2 + e S O 4 2 + S O 4
S 2 O 8 2 + h v 2 S O 4
S O 4 + H 2 O S O 4 2 + H + + O H
S O 4 + O H S O 4 2 + O H
S 2 O 8 2 + 2 H 2 O S O 4 2 + 3 H + + H O 2
S 2 O 8 2 + H O 2 S O 4 2 + H + + O 2
O 2 + O H O H + O 2 1
ZMIP OOH + S 2 O 8 2   ZMIP OO + H S O 4 + S O 4
ZMIP OH + S 2 O 8 2   ZMIP O + H S O 4 + S O 4
S 2 O 8 2 + S O 4 S O 4 2 + S 2 O 8
S 2 O 8 2 + O H O H + S 2 O 8
H S O 5 + e O H + S O 4
H S O 5 + h v O H + S O 4
S O 4 + H S O 5 S O 5 + H S O 4
S O 5 + H 2 O S O 4 2 + H + + O 2
ZMIP OOH + H S O 5   ZMIP OO + H 2 O + S O 4
ZMIP OH + H S O 5   ZMIP O + H 2 O + S O 4
H 2 O 2 + e O H + O H
H 2 O 2 + h v 2 O H

3. Materials and Methodology

3.1. Real Wastewater Samples

Table S2 lists the sampling locations of the raw wastewater specimens, while Table S3 tabulates their measured chemical and physical properties. Moreover, procedural details regarding wastewater sample preparation protocol are documented in Text S1.

3.2. Chemicals and Materials

Detailed information on the chemicals list and materials employed in this study is provided in Text S2 of the Supplementary Materials (SM).

3.3. Green Synthesis of the ZnO Nanoparticles

The comprehensive procedures for the preparation of water lettuce extract and the subsequent green synthesis of ZnO nanoparticles are described in Texts S3 and S4.

3.4. Construction of the MIP-202(Zr) Bio-MOF

The MIP-202(Zr) bio-MOF was synthesized via a mild hydrothermal method, as provided in Text S5 of the SM.

3.5. Fabrication of the ZMIP Nanocomposite

During the preparation of the MIP-202(Zr) bio-MOF, 5 g of the green-synthesized ZnO nanoparticles was introduced into the boiling flask immediately before initiating reflux heating, to achieve a 1:1 mass ratio of the two components. The subsequent synthesis steps were carried out in the same manner, and the resulting nanocomposite was designated as ZMIP.

3.6. Physicochemical Characterization of the Synthesized Materials

An outline of the analytical approaches employed to investigate the physicochemical characterizations of the fabricated catalysts is given in Text S6.

3.7. Point of Zero Charge

The point of zero charge (pHpzc) of the synthesized ZMIP nanocomposite was identified through the powder addition technique, commonly referred to as the pH drift method, as outlined in Text S7.

3.8. Optical Properties

The optical characteristics of the synthesized materials were comprehensively assessed following the procedures outlined in Text S8.

3.9. Experimental Set-Up and Degradation Study

An accurately weighed amount of CBZ was dissolved in deionized water to prepare a stock solution with a concentration of 50 mg/L. The photocatalytic degradation of CBZ was conducted in batch mode using a 250 mL glass reactor. Illumination was provided by a centrally mounted metal halide lamp (400 W) emitting visible light at λmax of 510 nm, with a 10 cm headspace left above the system. The reactor temperature was maintained at ambient value (25 ± 1 °C), and the solution pH was adjusted, as necessary, using 0.1 M solutions of sodium hydroxide (NaOH) and/or sulfuric acid (H2SO4). A digital magnetic stirrer hotplate (MSH-20D, DAIHAN Scientific Co., Ltd., Seoul, Republic of Korea) was used to agitate the reactor contents and regulate the reaction temperature. A multiparameter analyzer (HQ440d, HACH, Loveland, CO, USA) coupled with a pH sensor (PHC301, HACH, Loveland, CO, USA) was employed to measure and verify the solution pH. To initiate the photocatalytic reaction, the light source was switched on, and the pre-specified amount of the photocatalyst was added to 100 mL of pollutant solution. All experiments were conducted in triplicate to ensure the reproducibility of the results. A graphical overview of the experimental workflow is presented in Figure S6.

3.10. Experimental Design and Statistical Analysis

RSM was employed to assess the photocatalytic performance of the proposed system and to analyze the individual and interactive effects of the selected independent process variables on the dependent parameter or response (CBZ removal efficiency). Moreover, this methodology was utilized to optimize the independent variables based on maximizing the response as a target across three experimental levels coded as −1 (low), 0, and +1 (high). The independent parameters and their tested ranges were reaction time (D, 30–60 min), solution pH (F, 3–11), initial CBZ concentration (G, 15–25 mg/L), and photocatalyst dosage (H, 0.75–1.25 g/L), as presented in Table 5. The independent variables were incorporated into the Box–Behnken Design (BBD) for modeling. Including five replicated central points (Cp), a total of 29 experimental runs (Table 1) was determined using Equation (27) [128]. These experiments were employed to develop the design matrix and to predict the CBZ removal efficiency based on Equation (28) [129].
N = 2 F F 1 + C p
Y ( % ) = β 0 + i = 1 k β i   X i + i = 1 k β i i   X i 2   i = 1 k 1 j = i + 1 k β i j   X i j   + e    
where N represents the total number of experimental runs and F denotes the number of independent variables. The term Y (%) denotes the CBZ removal efficiency (response). Additionally, the β0, βi, βii, and βij terms represent the regression coefficients corresponding to the intercept, linear, quadratic, and interaction terms, respectively. Further, the coded independent variables are expressed as Xi and Xj (with i and j ranging from 1 to k), where k is the total number of independent variables (k = 4 in this study), and e corresponds to the model error (residual term). The experimental design was structured to provide a comprehensive assessment of variable interactions, thereby overcoming the inherent limitations of conventional one-factor-at-a-time (OFAT) approaches [130].
The Minitab® 22 statistical software was employed to perform regression analysis and to construct a polynomial model describing the overall process, thereby elucidating the relationship between the independent variables and response. Analysis of variance (ANOVA) was applied to evaluate the adequacy and reliability of the developed model [131]. The model’s validity was determined based on the correlation coefficient (R2), Fisher’s statistical values (F-values), and p-values, where a high F-value coupled with a p-value ≤ 0.05 indicated strong statistical significance in describing the relationship between the independent variables and the response [132].

3.11. Analytical Methods

The analytical methodologies applied throughout this investigation are thoroughly described and systematically outlined in Text S9 and Table S4.

3.12. Kinetic Study

In this work, the degradation kinetics of CBZ in the proposed system were evaluated by applying both pseudo-first-order and pseudo-second-order kinetic models to identify the model that most accurately characterizes the CBZ degradation process, as presented in Text S10.

4. Conclusions

This work reports the first successful integration of ZnO nanoparticles with the bio-MOF MIP-202(Zr) to construct a composite photocatalyst capable of efficiently degrading CBZ in real pharmaceutical wastewater under visible-light irradiation. The characterization analyses verified the successful fabrication and interaction between Zn and the MIP-202(Zr) framework. Under identical conditions (pH = 7, catalyst dosage = 1 g/L, initial CBZ concentration = 20 mg/L, and reaction time = 60 min), the photocatalytic efficiency of the ZMIP nanocomposite reached 76.59%, which was significantly higher than that of pure ZnO and MIP-202(Zr) due to the reduction in bandgap in ZMIP to 2.74 ± 0.1 eV and the improvement of charge carrier separation. The optimum conditions were identified using RSM as a reaction time of 90 min, solution pH of 6, initial CBZ concentration of 15 mg/L, and a photocatalyst dosage of 1.25 g/L, under which CBZ removal achieved 99.37% with a corresponding TOC mineralization of 84.39%. The ZMIP photocatalyst exhibited high stability and reusability, maintaining CBZ degradation efficiencies of 99.37, 96.80, 94.54, 85.23, and 77.97% across five successive cycles, thereby confirming its sustained catalytic performance during repeated operation. Under visible-light irradiation, the ZMIP heterojunction promotes efficient charge separation by transferring electrons from ZnO to MIP-202(Zr) and holes in the opposite direction. This process suppresses e/h+ recombination and facilitates the generation of ROS (OH and O2• − radicals), which synergistically drive CBZ oxidation. Further, the quenching experiments identified OH radicals as the dominant reactive species in the CBZ degradation process, followed by O2• − radicals, while the photogenerated holes played the least significant role. On the other hand, the degradation pathways of CBZ were elucidated through the identification of transformation products using UHPLC-MS analysis. Furthermore, under the optimum operating conditions, the ZMIP/light photocatalytic system demonstrated strong catalytic activity toward a range of pharmaceutical contaminants (15 mg/L), achieving removal efficiencies of 82.93% for doxorubicin hydrochloride, 76.84% for tetracycline, 72.08% for paracetamol, and 67.71% for ibuprofen, in comparison with 99.37% for CBZ. These findings highlight the broad applicability of the ZMIP/light system, demonstrating its potential as a versatile photocatalyst for the effective removal of structurally diverse pharmaceutical pollutants. In addition, the system maintained high CBZ degradation performance across different water matrices, with efficiencies of 92.81% in tap water, 70.86% in seawater, 54.06% in lake water, and 31.08% in drain water, compared to 99.37% in deionized water. Further, the degradation system exhibited lower performance in the presence of high NOM levels and inorganic salts; therefore, these compounds should be removed or minimized before applying the ZMIP/light system in real wastewater treatment to achieve optimal photocatalytic performance. Significantly, the ZMIP/light system demonstrated consistent and robust performance in the treatment of real industrial wastewater obtained from a pharmaceutical plant. The substantial TOC removal achieved (78.37% after 90 min) highlights the practical applicability of the ZMIP/light system, emphasizing its potential as a reliable photocatalytic technology for the advanced treatment of complex industrial wastewater streams. In addition, the introduction of highly reactive inorganic oxidants markedly enhanced CBZ degradation in the ZMIP/light system. Among the oxidants tested, KIO4 exhibited the greatest enhancement, achieving 98.93% removal, followed by K2S2O8 (92.15%), KHSO5 (89.22%), and H2O2 (81.86%), compared with 76.59% in the absence of additives. This enhancement indicates that integrating the ZMIP/light system with suitable inorganic oxidants, particularly KIO4, can significantly improve photocatalytic efficiency, providing a promising approach for the near-complete elimination of pharmaceutical pollutants. In the future, the economic feasibility and environmental impacts of the proposed degradation system can be investigated. Furthermore, to advance toward practical implementation, the system should be evaluated using real wastewater matrices, alongside the design and optimization of large-scale photocatalytic reactors.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/catal15111017/s1, Text S1: Processing and conditioning procedures of raw wastewater samples; Text S2: Chemicals and materials; Text S3: Preparation of the water lettuce plant extract; Text S4: Green synthesis of ZnO nanoparticles; Text S5: Fabrication of the MIP-202(Zr) bio-MOF; Text S6: Characterization of the developed compounds; Text S7: Point of zero charge; Text S8: Optical properties; Text S9: Analytical methods; Text S10: Kinetic study; Figure S1: Combined XRD patterns of the synthesized materials; Figure S2: Particle size distribution of the green-synthesized ZnO nanoparticles; Figure S3: Particle size distribution of the MIP-202(Zr) bio-MOF; Figure S4: Point of zero charge of the ZMIP photocatalyst; Figure S5: m/z peaks of CBZ and its generated by-products; Figure S6: Graphical outline of the experimental procedures; Table S1: Comparison of the main diffraction peaks of MIP-202(Zr) with reported literature values; Table S2: Sources of raw wastewater samples; Table S3: Characterization properties of distinct water matrices; Table S4: Wavelengths for the UV-Vis spectrophotometric detection of various pollutants; references [31,34,74,77,79,92,96,97,133,134,135,136,137,138,139,140,141,142,143,144,145,146] are cited in the Supplementary Materials.

Author Contributions

M.M.G.: Conceptualization, Data curation, Formal analysis, Investigation, Methodology, Validation, Visualization, Writing—original draft, and Writing—review and editing. A.T.: Data curation, Formal analysis, Funding acquisition, Investigation, Project administration, Resources, Software, Supervision, Validation, Visualization, and Writing—review and editing. A.M.E.: Data curation, Formal analysis, Investigation, Methodology, Software, Validation, Visualization, and Writing—review and editing. S.A.A.-H.: Data curation, Formal analysis, Investigation, Methodology, Validation, Visualization, and Writing—review and editing. E.M.M.: Formal analysis, Investigation, Software, Supervision, Validation, and Writing—review and editing. H.S.: Data curation, Formal analysis, Investigation, Methodology, Validation, Visualization, and Writing—original draft. M.S.: Data curation, Formal analysis, Investigation, Methodology, Supervision, Validation, Visualization, and Writing—review and editing. M.E.: Conceptualization, Investigation, Methodology, Supervision, Validation, and Writing—review and editing. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported and funded by the Deanship of Scientific Research at Imam Mohammad Ibn Saud Islamic University (IMSIU) (grant number IMSIU-DDRSP2502).

Data Availability Statement

All data will be made available as published and as requested through the corresponding authors’ contacts.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Bakhadur, A.; Alem, H.; Carré, V.; Gries, T.; Balan, L.; Cantin, J.L.; Medjahdi, G.; Orazov, Z.; Uralbekov, B.; Schneider, R. Porous Nitrogen-Doped TiO2/Graphene Oxide Derived from H2N-MIL-125(Ti)/Graphene Oxide Composites as Highly Efficient Visible Light Active Photocatalysts for the Degradation of Dyes and Carbamazepine. Appl. Surf. Sci. 2025, 714, 164466. [Google Scholar] [CrossRef]
  2. Mohammed-Amine, E.; Kaltoum, B.; El Mountassir, E.M.; Abdelaziz, A.T.; Stephanie, R.; Stephanie, L.; Anne, P.; Pascal, W.W.C.; Alrashed, M.M.; Salah, R. Novel Sol-Gel Synthesis of TiO2/BiPO4 Composite for Enhanced Photocatalytic Degradation of Carbamazepine under UV and Visible Light: Kinetic, Identification of Photoproducts and Mechanistic Insights. J. Water Process Eng. 2025, 70, 107098. [Google Scholar] [CrossRef]
  3. Chong, M.N.; Jin, B. Photocatalytic Treatment of High Concentration Carbamazepine in Synthetic Hospital Wastewater. J. Hazard Mater. 2012, 199, 135–142. [Google Scholar] [CrossRef]
  4. Hai, F.I.; Yang, S.; Asif, M.B.; Sencadas, V.; Shawkat, S.; Sanderson-Smith, M.; Gorman, J.; Xu, Z.Q.; Yamamoto, K. Carbamazepine as a Possible Anthropogenic Marker in Water: Occurrences, Toxicological Effects, Regulations and Removal by Wastewater Treatment Technologies. Water 2018, 10, 107. [Google Scholar] [CrossRef]
  5. Novikov, M.V.; Snytnikova, O.A.; Fedunov, R.G.; Yanshole, V.V.; Grivin, V.P.; Plyusnin, V.F.; Xu, J.; Pozdnyakov, I.P. A New View on the Mechanism of UV Photodegradation of the Tricyclic Antidepressant Carbamazepine in Aqueous Solutions. Chemosphere 2023, 329, 138652. [Google Scholar] [CrossRef] [PubMed]
  6. Samy, M.; Ibrahim, M.G.; Fujii, M.; Diab, K.E.; ElKady, M.; Gar Alalm, M. CNTs/MOF-808 Painted Plates for Extended Treatment of Pharmaceutical and Agrochemical Wastewaters in a Novel Photocatalytic Reactor. Chem. Eng. J. 2021, 406, 127152. [Google Scholar] [CrossRef]
  7. He, Y.; Ding, L.; Zhou, J.; Liu, D.; Zhang, Y. Synergistic Activation of Photocatalysis and Peroxymonosulfate via Fe2O3/Bi2O3 p-n Heterojunction for Enhanced Degradation of Tetracycline. Chem. Phys. Lett. 2025, 876, 142301. [Google Scholar] [CrossRef]
  8. Li, H.; Wang, K.; Xu, J.; Li, T.; Lu, Y.; Zou, R.; Zheng, G. Pre-Reduction of Goethite by Hydroxylamine Hydrochloride Enhances Sulfamethoxazole Degradation via Accelerated Fe(III)/Fe(II) Cycling in Fenton-like Systems. Chem. Eng. J. 2025, 522, 167877. [Google Scholar] [CrossRef]
  9. Montenegro-Apraez, D.; Graça, C.A.L.; Machuca-Martinez, F.; Soares, O.S.G.P. Catalytic Ozonation of Lamotrigine with Modified Activated Carbon: Degradation, Mineralization and by-Products Analysis. J. Water Process Eng. 2025, 77, 108434. [Google Scholar] [CrossRef]
  10. Butola, D.; Purohit, L.P. Synergistic G-C3N4@ZnO/SnO2 Heterojunction Nanocomposites for Multifunctional Applications in Photocatalysis and Gas Sensing. Surf. Interfaces 2025, 72, 106963. [Google Scholar] [CrossRef]
  11. Quan, Y.; Lu, S.; Wang, Q.; Wang, H.; Hu, E.; Xin, X.; Su, Y.; Zhang, Y.; Bao, J. Elucidating Structure-Activity Relationships in CdS/ZnO Heterojunctions for Synergistic Adsorption-Photocatalysis of Uranium (VI) Removal. Sep. Purif. Technol. 2025, 375, 133806. [Google Scholar] [CrossRef]
  12. Sonia; Kumari, H.; Chahal, S.; Suman; Kumar, S.; Mahak; Kumar, P.; Kumar, A. Hydrothermally Synthesized ZnFe2O4/ZnO Heterojunction Nanocomposites for Enhanced RB Dye Degradation via Z-Scheme Photocatalysis. Mater. Chem. Phys. 2024, 322, 129560. [Google Scholar] [CrossRef]
  13. Jayathilake, K.M.P.I.; Manage, P.M.; Idroos, F.S. Development of Water Lettuce (Pistia Spp.) Based Biochar Filter for the Treatment of Industrial Wastewater: A Green Approach. Curr. Sci. 2024, 127, 1208–1218. [Google Scholar] [CrossRef]
  14. Thongam, D.D.; Hang, D.R.; Gupta, J. Highly Active Sunlight-Driven Photocatalysis: Harnessing Z-Scheme Charge Transfer in g-C3N4/ZnO Heterojunction for Effective Advanced Oxidation Processes. Surf. Interfaces 2025, 72, 107048. [Google Scholar] [CrossRef]
  15. Qin, B.; Chen, J.; Xu, F.; Zou, J.; Wang, W.; Lv, Y.; Tian, L.; Zhou, L.; Feng, J. ZIF-8 Derived ZnS/ZnO Type-II Heterojunction for Efficient Photocatalytic Hydrogen Peroxide Production and Norfloxacin Degradation. J. Alloys Compd. 2025, 1038, 182866. [Google Scholar] [CrossRef]
  16. Oliveira, R.A.; Castro, M.A.M.; Porto, D.L.; Aragão, C.F.S.; Souza, R.P.; Silva, U.C.; Bomio, M.R.D.; Motta, F.V. Immobilization of Bi2MoO6/ZnO Heterojunctions on Glass Substrate: Design of Drug and Dye Mixture Degradation by Solar-Driven Photocatalysis. J. Photochem. Photobiol. A Chem. 2024, 452, 115619. [Google Scholar] [CrossRef]
  17. Luo, H.; Feng, M.; Zheng, S.; Wang, W.; Li, X. Enhancing Sunlight-Driven Photocatalysis of Ag/AgCl@MOF-808 Heterojunction via Synergistic Effects of Plasmonic Ag QDs and Heterostructure. J. Environ. Chem. Eng. 2025, 13, 117456. [Google Scholar] [CrossRef]
  18. Berehe, B.A.; Desalew, A.A.; Derbe, G.W.; Misganaw, D.M.; Mohammed, K.S.; Chang, J.Y.; Girma, W.M. Enhanced Photocatalytic Degradation of Methylene Blue Dye via Valorization of a Polyethylene Terephthalate Plastic Waste-Derived Metal–Organic Framework-Based ZnO@Co-BDC Composite Catalyst. Nanoscale Adv. 2025, 7, 3834–3845. [Google Scholar] [CrossRef] [PubMed]
  19. Hu, Z.; Liao, J.; Zhou, J.; Zhao, L.; Liu, Y.; Zhang, Y.; Chen, W.; Tang, S. A New Green Approach to Synthesizing MIP-202@porous Silica Microspheres for Positional Isomer/Enantiomer/Hydrophilic Separation. Chin. Chem. Lett. 2025, 36, 109985. [Google Scholar] [CrossRef]
  20. Sandhu, S.S.; Kotagiri, Y.G.; Ashvin Iresh Fernando, I.P.U.; Kalaj, M.; Tostado, N.; Teymourian, H.; Alberts, E.M.; Thornell, T.L.; Jenness, G.R.; Harvey, S.P.; et al. Green MIP-202(Zr) Catalyst: Degradation and Thermally Robust Biomimetic Sensing of Nerve Agents. J. Am. Chem. Soc. 2021, 143, 18261–18271. [Google Scholar] [CrossRef]
  21. Roushree, R.R.; Haimbodi, R. Recent Advances in ZnO/MOFs: Synthesis, Characterization, and Applications in Sustainable Antibiotic Wastewater Treatment. Chin. J. Anal. Chem. 2025, 100619. [Google Scholar] [CrossRef]
  22. Zhang, Z.; Bai, Z.; Yu, S.; Meng, X.; Xiao, S. Photo-Fenton Efficient Degradation of Organic Pollutants over S-Scheme ZnO@NH2-MIL-88B Heterojunction Established for Electron Transfer Channel. Chem. Eng. Sci. 2024, 288, 119789. [Google Scholar] [CrossRef]
  23. Fu, S.; Xi, W.; Ren, J.; Wei, H.; Sun, W. Study on the Photocatalytic Properties of Metal–Organic Framework-Derived C-, N-Co-Doped ZnO. Materials 2024, 17, 855. [Google Scholar] [CrossRef]
  24. Jin, C.; Ge, C.; Jian, Z.; Wei, Y. Facile Synthesis and High Photocatalytic Degradation Performance of ZnO-SnO2 Hollow Spheres. Nanoscale Res. Lett. 2016, 11, 526. [Google Scholar] [CrossRef]
  25. Piri, A.; Kaykhaii, M.; Khajeh, M.; Oveisi, A.R. Application of a Magnetically Separable Zr-MOF for Fast Extraction of Palladium before Its Spectrophotometric Detection. BMC Chem. 2024, 18, 63. [Google Scholar] [CrossRef]
  26. Abdelghani, G.M.; Ahmed, A.B.; Al-Zubaidi, A.B. Synthesis, Characterization, and the Influence of Energy of Irradiation on Optical Properties of ZnO Nanostructures. Sci. Rep. 2022, 12, 20016. [Google Scholar] [CrossRef] [PubMed]
  27. Murthy, K.R.S.; Raghu, G.K.; Binnal, P. Zinc Oxide Nanostructured Material for Sensor Application. J. Biotechnol. Bioeng. 2021, 5, 25–29. [Google Scholar] [CrossRef]
  28. Flores-Loyola, D.M.; Márquez-Guerrero, E.; Galindo-Guzman, S.Y.; Marszalek, M.; Green, J.E.; Monserrat Sánchez-Pérez, D.; Flores-Loyola, E.; Yuridia Márquez-Guerrero, S.; Galindo-Guzman, M.; Marszalek, J.E. Green Synthesis and Characterization of Zinc Oxide Nanoparticles Using Larrea tridentata Extract and Their Impact on the In-Vitro Germination and Seedling Growth of Capsicum annuum. Sustainability 2023, 15, 3080. [Google Scholar] [CrossRef]
  29. Ramesh, P.; Saravanan, K.; Manogar, P.; Johnson, J.; Vinoth, E.; Mayakannan, M. Green Synthesis and Characterization of Biocompatible Zinc Oxide Nanoparticles and Evaluation of Its Antibacterial Potential. Sens. Biosens. Res. 2021, 31, 100399. [Google Scholar] [CrossRef]
  30. Hashem, T.; Ibrahim, A.H.; Wöll, C.; Alkordi, M.H. Grafting Zirconium-Based Metal-Organic Framework UiO-66-NH2 Nanoparticles on Cellulose Fibers for the Removal of Cr(VI) Ions and Methyl Orange from Water. ACS Appl. Nano Mater. 2019, 2, 5804–5808. [Google Scholar] [CrossRef]
  31. Bagherzadeh, M.; Chegeni, M.; Bayrami, A.; Amini, M. Superior and Efficient Performance of Cost-Effective MIP-202 Catalyst over UiO-66-(CO2H)2 in Epoxide Ring Opening Reactions. Sci. Rep. 2024, 14, 17730. [Google Scholar] [CrossRef] [PubMed]
  32. Han, G.; Qian, Q.; Mizrahi Rodriguez, K.; Smith, Z.P. Hydrothermal Synthesis of Sub-20 Nm Amine-Functionalized MIL-101(Cr) Nanoparticles with High Surface Area and Enhanced CO2Uptake. Ind. Eng. Chem. Res. 2020, 59, 7888–7900. [Google Scholar] [CrossRef]
  33. Xu, X.; Xia, L.; Zheng, C.; Liu, Y.; Yu, D.; Li, J.; Zhong, S.; Li, C.; Song, H.; Liu, Y.; et al. Unravelling Nonclassical Beam Damage Mechanisms in Metal-Organic Frameworks by Low-Dose Electron Microscopy. Nat. Commun. 2025, 16, 261. [Google Scholar] [CrossRef] [PubMed]
  34. Gaber, M.M.; Shokry, H.; Samy, M.; El-Bestawy, E.A. Green Approach for Fabricating Hybrids of Food Waste-Derived Biochar/Zinc Oxide for Effective Degradation of Bromothymol Blue Dye in a Photocatalysis/Persulfate Activation System. Chemosphere 2024, 364, 143245. [Google Scholar] [CrossRef] [PubMed]
  35. da Trindade, L.G.; Borba, K.M.N.; Trench, A.B.; Zanchet, L.; Teodoro, V.; Pontes, F.M.L.; Longo, E.; Mazzo, T.M. Effective Strategy to Coupling Zr-MOF/ZnO: Synthesis, Morphology and Photoelectrochemical Properties Evaluation. J. Solid. State Chem. 2021, 293, 121794. [Google Scholar] [CrossRef]
  36. Taddei, M.; Schukraft, G.M.; Warwick, M.E.A.; Tiana, D.; McPherson, M.J.; Jones, D.R.; Petit, C. Band Gap Modulation in Zirconium-Based Metal–Organic Frameworks by Defect Engineering. J. Mater. Chem. A Mater. 2019, 7, 23781–23786. [Google Scholar] [CrossRef]
  37. Parsa Amouzesh, S.; Zandjou, M.; Ali Khodadadi, A.; Mortazavi, Y.; Hooriabad Saboor, F.; Saris, S.; Javanmard, A.; Alirezayi, S.; Asgari, M. Innovative Photocatalyst Design: Advancing ZnO/MIL-100(Fe) through Atomic Layer Deposition in Hydrogen Evolution. ChemCatChem 2024, 16, e202401016. [Google Scholar] [CrossRef]
  38. Harisankar, A.; Preethi, P.C.; Sreeja, T.G.; Rejani, P.; Murali, M.; Raghunandan, R. Zinc Oxide Functionalized MOF-5 for the Adsorptive Removal of Pb(II) Metal Ions and Photocatalytic Degradation of Methylene Blue Dye in Aqueous Medium. Ionics 2024, 30, 2313–2331. [Google Scholar] [CrossRef]
  39. Wang, J.; Bi, L.; Fu, Q.; Jen, A.K.Y. Methods for Passivating Defects of Perovskite for Inverted Perovskite Solar Cells and Modules. Adv. Energy Mater. 2024, 14, 2401414. [Google Scholar] [CrossRef]
  40. Ul Muazzam, U.; Muralidharan, R.; Raghavan, S.; Nath, D.N. Investigation of Optical Functions, Sub-Bandgap Transitions, and Urbach Tail in the Absorption Spectra of Ga2O3 Thin Films Deposited Using Mist-CVD. Opt. Mater. 2023, 145, 114373. [Google Scholar] [CrossRef]
  41. Vetokhina, V.; Nepomniashchaia, N.; de Prado, E.; Pacherova, O.; Kocourek, T.; Anandakrishnan, S.S.; Bai, Y.; Dejneka, A.; Tyunina, M. Tuning Optical Absorption in Perovskite (K,Na)NbO3 Ferroelectrics. Mater. Adv. 2024, 5, 8901–8908. [Google Scholar] [CrossRef]
  42. Ganesha Krishna, V.S.; Mahesha, M.G. (Mg,Mn)-Dual Doping Synergism towards Luminescence and Electrical Properties of ZnO/p-Si Heterojunction Diodes. RSC Adv. 2023, 13, 32282–32295. [Google Scholar] [CrossRef]
  43. Klein, J.; Kampermann, L.; Mockenhaupt, B.; Behrens, M.; Strunk, J.; Bacher, G.; Klein, J.; Kampermann, L.; Bacher, G.; Mockenhaupt, B.; et al. Limitations of the Tauc Plot Method. Adv. Funct. Mater. 2023, 33, 2304523. [Google Scholar] [CrossRef]
  44. Sohrabnezhad, S.; Pourahmad, A.; Karimi, M.F. Magnetite-Metal Organic Framework Core@shell for Degradation of Ampicillin Antibiotic in Aqueous Solution. J. Solid. State Chem. 2020, 288, 121420. [Google Scholar] [CrossRef]
  45. Zhao, W.; Yan, M.; Chen, Y.; Shen, J.; Hong, X.; Mu, F.; Li, S.; Zhang, S.; Wang, Q.; Dai, B.; et al. Rational Design of Novel Metal-Organic Framework/Bi4O7 S-Scheme Heterojunction Photocatalyst for Boosting Carbamazepine Degradation. Appl. Surf. Sci. 2023, 622, 156876. [Google Scholar] [CrossRef]
  46. Goh, J.W.; Xiong, Y.; Wu, W.; Huang, Z.; Ong, S.L.; Hu, J.Y. Degradation of Carbamazepine by HF-Free-Synthesized MIL-101(Cr)@Anatase TiO2 Composite under UV-A Irradiation: Degradation Mechanism, Wastewater Matrix Effect, and Degradation Pathway. Water 2022, 14, 3964. [Google Scholar] [CrossRef]
  47. Liu, N.; Shang, Q.; Gao, K.; Cheng, Q.; Pan, Z. Construction of ZnO/ZIF-9 Heterojunction Photocatalyst: Enhanced Photocatalytic Performance and Mechanistic Insight. New J. Chem. 2020, 44, 6384–6393. [Google Scholar] [CrossRef]
  48. Du, Q.; Wu, P.; Sun, Y.; Zhang, J.; He, H. Selective Photodegradation of Tetracycline by Molecularly Imprinted ZnO@NH2-UiO-66 Composites. Chem. Eng. J. 2020, 390, 124614. [Google Scholar] [CrossRef]
  49. Samy, M.; Ibrahim, M.G.; Gar Alalm, M.; Fujii, M. MIL-53(Al)/ZnO Coated Plates with High Photocatalytic Activity for Extended Degradation of Trimethoprim via Novel Photocatalytic Reactor. Sep. Purif. Technol. 2020, 249, 117173. [Google Scholar] [CrossRef]
  50. Mohsen Mousavi, S.; Chamack, M.; Fakhri, H. Study of New Hybrid Material of ZnO/CuO and Metal-Organic Framework as Photocatalyst for Removal of Tetracycline from Water. J. Nanostruct 2022, 12, 1097–1107. [Google Scholar] [CrossRef]
  51. Fakhri, H.; Bagheri, H. Highly Efficient Zr-MOF@WO3/Graphene Oxide Photocatalyst: Synthesis, Characterization and Photodegradation of Tetracycline and Malathion. Mater. Sci. Semicond. Process 2020, 107, 104815. [Google Scholar] [CrossRef]
  52. Chen, J.; Zhang, X.; Shi, X.; Bi, F.; Yang, Y.; Wang, Y. Synergistic Effects of Octahedral TiO2-MIL-101(Cr) with Two Heterojunctions for Enhancing Visible-Light Photocatalytic Degradation of Liquid Tetracycline and Gaseous Toluene. J. Colloid. Interface Sci. 2020, 579, 37–49. [Google Scholar] [CrossRef]
  53. Wu, J.; Fang, X.; Zhu, Y.; Ma, N.; Dai, W. Well-Designed TiO2@UiO-66-NH2Nanocomposite with Superior Photocatalytic Activity for Tetracycline under Restricted Space. Energy Fuels 2020, 34, 12911–12917. [Google Scholar] [CrossRef]
  54. Hu, P.; Yao, C.; Yang, L.; Xin, Y.; Miao, Y. Boosted Photodegradation of Tetracycline Hydrochloride over Z-Scheme MIL-88B(Fe)/Bi2WO6 Composites under Visible Light. Colloids Surf. A Physicochem. Eng. Asp. 2021, 627, 127248. [Google Scholar] [CrossRef]
  55. Yin, S.; Chen, Y.; Li, M.; Hu, Q.; Ding, Y.; Shao, Y.; Di, J.; Xia, J.; Li, H. Construction of NH2-MIL-125(Ti)/Bi2WO6 Composites with Accelerated Charge Separation for Degradation of Organic Contaminants under Visible Light Irradiation. Green. Energy Environ. 2020, 5, 203–213. [Google Scholar] [CrossRef]
  56. Mirzaei, A.; Chen, Z.; Haghighat, F.; Yerushalmi, L. Removal of Pharmaceuticals and Endocrine Disrupting Compounds from Water by Zinc Oxide-Based Photocatalytic Degradation: A Review. Sustain. Cities Soc. 2016, 27, 407–418. [Google Scholar] [CrossRef]
  57. Mishra, S.R.; Gadore, V.; Ahmaruzzaman, M. Insights into Persulfate-Activated Photodegradation of Tinidazole and Photoreduction of Hexavalent Chromium through β-In2S3 Anchored on Ag-Doped Fish Scale-Derived HAp Composite Quantum Dots. J. Clean. Prod. 2023, 427, 139221. [Google Scholar] [CrossRef]
  58. Dehghan, A.; Dehghani, M.H.; Nabizadeh, R.; Ramezanian, N.; Alimohammadi, M.; Najafpoor, A.A. Adsorption and Visible-Light Photocatalytic Degradation of Tetracycline Hydrochloride from Aqueous Solutions Using 3D Hierarchical Mesoporous BiOI: Synthesis and Characterization, Process Optimization, Adsorption and Degradation Modeling. Chem. Eng. Res. Des. 2018, 129, 217–230. [Google Scholar] [CrossRef]
  59. Gar Alalm, M.; Samy, M.; Ookawara, S.; Ohno, T. Immobilization of S-TiO2 on Reusable Aluminum Plates by Polysiloxane for Photocatalytic Degradation of 2,4-Dichlorophenol in Water. J. Water Process Eng. 2018, 26, 329–335. [Google Scholar] [CrossRef]
  60. Liu, X.; Lv, P.; Yao, G.; Ma, C.; Huo, P.; Yan, Y. Microwave-Assisted Synthesis of Selective Degradation Photocatalyst by Surface Molecular Imprinting Method for the Degradation of Tetracycline onto ClTiO2. Chem. Eng. J. 2013, 217, 398–406. [Google Scholar] [CrossRef]
  61. Fathinia, M.; Khataee, A.R. Residence Time Distribution Analysis and Optimization of Photocatalysis of Phenazopyridine Using Immobilized TiO2 Nanoparticles in a Rectangular Photoreactor. J. Ind. Eng. Chem. 2013, 19, 1525–1534. [Google Scholar] [CrossRef]
  62. Zhang, X.; Zhang, X.; Li, H.; Ao, X.; Sun, W.; Li, Z. Reactive Oxygen Species Generated in Situ During Carbamazepine Photodegradation at 222 Nm Far-UVC: Unexpected Role of H2O Molecules. Environ. Sci. Technol. 2024, 58, 19070–19079. [Google Scholar] [CrossRef]
  63. Zayyat, R.M.; Yahfoufi, R.; Al-Hindi, M.; Kordahi, M.A.; Ayoub, G.M.; Ahmad, M.N. Elucidating the Dynamics of Carbamazepine Uptake Using Date Pit-Derived Activated Carbon: A Comprehensive Kinetic and Thermodynamic Analysis. Heliyon 2024, 10, e39068. [Google Scholar] [CrossRef]
  64. Abumelha, H.M.; Alzahrani, S.O.; Alrefaee, S.H.; Al-bonayan, A.M.; Alkhatib, F.; Saad, F.A.; El-Metwaly, N.M. Evaluation of Tetracycline Removal by Magnetic Metal Organic Framework from Aqueous Solutions: Adsorption Isotherm, Kinetics, Thermodynamics, and Box-Behnken Design Optimization. J. Saudi Chem. Soc. 2023, 27, 101706. [Google Scholar] [CrossRef]
  65. Ahmad, F.A. The Use of Agro-Waste-Based Adsorbents as Sustainable, Renewable, and Low-Cost Alternatives for the Removal of Ibuprofen and Carbamazepine from Water. Heliyon 2023, 9, e16449. [Google Scholar] [CrossRef]
  66. Kumar, S.; Dhiman, V.; Kumar, R.; Kaur, S.; Sharma, P.; Singh, K. Unveiling the Photocatalytic Properties of Cadmium Oxide for Sustainable Approach towards Water Remediation: A Review. Coord. Chem. Rev. 2025, 539, 216713. [Google Scholar] [CrossRef]
  67. Li, Z.; Tao, W.; Wang, Y.; Ye, X.; Chen, Y.; Han, B.; Lee, L.Y.S. Corrosion-Resistant MoO3/Fe2O3/MoS2 Heterojunctions Stabilize OH- Adsorption for Efficient Light-Assisted Seawater Electrooxidation. J. Am. Chem. Soc. 2025, 147, 24461. [Google Scholar] [CrossRef]
  68. Décima, M.A.; Marzeddu, S.; Barchiesi, M.; Di Marcantonio, C.; Chiavola, A.; Boni, M.R. A Review on the Removal of Carbamazepine from Aqueous Solution by Using Activated Carbon and Biochar. Sustainability 2021, 13, 11760. [Google Scholar] [CrossRef]
  69. Sultana, M.; Mohapatra, S.R.; Ahmaruzzaman, M. Spherical Magnetic MgO-CeO2-Fe3O4@JB Heterojunction for Enhanced Photodegradation of Pesticide and Complex Anionic Dyes: Understanding the Degradation Process and Diverse Water Systems. Chem. Eng. J. 2024, 502, 157549. [Google Scholar] [CrossRef]
  70. Mohtaram, M.S.; Sabbaghi, S.; Rasouli, J.; Rasouli, K. Photocatalytic Degradation of Tetracycline Using a Novel WO3–ZnO/AC under Visible Light Irradiation: Optimization of Effective Factors by RSM-CCD. Environ. Pollut. 2024, 347, 123746. [Google Scholar] [CrossRef] [PubMed]
  71. Quy, B.M.; Thu, N.T.N.; Xuan, V.T.; Hoa, N.T.H.; Linh, N.T.N.; Tung, V.Q.; Le, V.T.T.; Thao, T.T.; Ngan, N.T.K.; Tho, P.T.; et al. Photocatalytic Degradation Performance of a Chitosan/ZnO–Fe3O4 Nanocomposite over Cationic and Anionic Dyes under Visible-Light Irradiation. RSC Adv. 2025, 15, 1590–1603. [Google Scholar] [CrossRef] [PubMed]
  72. Tran, H.D.; Nguyen, D.Q.; Do, P.T.; Tran, U.N.P. Kinetics of Photocatalytic Degradation of Organic Compounds: A Mini-Review and New Approach. RSC Adv. 2023, 13, 16915–16925. [Google Scholar] [CrossRef]
  73. Bloh, J.Z. A Holistic Approach to Model the Kinetics of Photocatalytic Reactions. Front. Chem. 2019, 7, 440665. [Google Scholar] [CrossRef] [PubMed]
  74. Gaber, M.M.; Toghan, A.; Shokry, H.; Samy, M. Efficient Oxidative Degradation of Organic Pollutants in Real Industrial Effluents Using a Green-Synthesized Magnetite Supported on Biochar Catalyst. RSC Adv. 2025, 15, 31522–31538. [Google Scholar] [CrossRef]
  75. Gaber, M.M.; Rashid, N.; Alzahrani, A.; Alanazi, F. Utilization of Incense Stick Residues as a Sustainable Catalyst for Efficient Peroxydisulfate Activation in the Oxidative Degradation of Atrazine from Real Industrial Effluents. J. Environ. Chem. Eng. 2025, 13, 116031. [Google Scholar] [CrossRef]
  76. Abdel-Salam, M.O.; Farghal, H.H.; El Sawy, E.; Yoon, T.; El-Sayed, M.M.H. Activation of Peroxymonosulfate for Rhodamine-B Removal from Water: Enhanced Efficiency with Cobalt-Enriched, Magnetically Recoverable CNTs. RSC Adv. 2025, 15, 6371–6383. [Google Scholar] [CrossRef]
  77. El-Bestawy, E.A.; Gaber, M.; Shokry, H.; Samy, M. Effective Degradation of Atrazine by Spinach-Derived Biochar via Persulfate Activation System: Process Optimization, Mechanism, Degradation Pathway and Application in Real Wastewater. Environ. Res. 2023, 229, 115987. [Google Scholar] [CrossRef]
  78. Mathew, R.A.; Kanmani, S. Photocatalytic Degradation of Carbamazepine Using Ozonation and Photocatalytic Ozonation with TiO2 and WO3. Water Pr. Technol. 2020, 15, 645–651. [Google Scholar] [CrossRef]
  79. Gaber, M.M.; Samy, M.; El-Bestawy, E.A.; Shokry, H. Effective Degradation of Tetracycline and Real Pharmaceutical Wastewater Using Novel Nanocomposites of Biosynthesized ZnO and Carbonized Toner Powder. Chemosphere 2024, 352, 141448. [Google Scholar] [CrossRef] [PubMed]
  80. Franz, S.; Falletta, E.; Arab, H.; Murgolo, S.; Bestetti, M.; Mascolo, G. Degradation of Carbamazepine by Photo(Electro)Catalysis on Nanostructured TiO2 Meshes: Transformation Products and Reaction Pathways. Catalysts 2020, 10, 169. [Google Scholar] [CrossRef]
  81. Grinnell, C.; Samokhvalov, A. Exploring the Electronic Structure of Aluminum Metal-Organic Framework Basolite A100: Solid-State Synchronous Fluorescence Spectroscopy Reveals New Charge Excitation/Relaxation Pathways. Phys. Chem. Chem. Phys. 2018, 20, 26947–26956. [Google Scholar] [CrossRef]
  82. Kumar, A.; Kumar, A.; Sharma, G.; Al-Muhtaseb, A.H.; Naushad, M.; Ghfar, A.A.; Stadler, F.J. Quaternary Magnetic BiOCl/g-C3N4/Cu2O/Fe3O4 Nano-Junction for Visible Light and Solar Powered Degradation of Sulfamethoxazole from Aqueous Environment. Chem. Eng. J. 2018, 334, 462–478. [Google Scholar] [CrossRef]
  83. Ajmal, A.; Majeed, I.; Malik, R.N.; Idriss, H.; Nadeem, M.A. Principles and Mechanisms of Photocatalytic Dye Degradation on TiO2 Based Photocatalysts: A Comparative Overview. RSC Adv. 2014, 4, 37003–37026. [Google Scholar] [CrossRef]
  84. Zhou, C.; Lai, C.; Xu, P.; Zeng, G.; Huang, D.; Li, Z.; Zhang, C.; Cheng, M.; Hu, L.; Wan, J.; et al. Rational Design of Carbon-Doped Carbon Nitride/Bi12O17Cl2 Composites: A Promising Candidate Photocatalyst for Boosting Visible-Light-Driven Photocatalytic Degradation of Tetracycline. ACS Sustain. Chem. Eng. 2018, 6, 6941–6949. [Google Scholar] [CrossRef]
  85. Samy, M.; Ibrahim, M.G.; Gar Alalm, M.; Fujii, M. Effective Photocatalytic Degradation of Sulfamethazine by CNTs/LaVO4 in Suspension and Dip Coating Modes. Sep. Purif. Technol. 2020, 235, 116138. [Google Scholar] [CrossRef]
  86. Zeng, H.; Chen, Y.; Xu, J.; Li, S.; Wu, J.; Li, D.; Zhang, J. Iron-Based Materials for Activation of Periodate in Water and Wastewater Treatment Processes: The Important Role of Fe Species. Chem. Eng. J. 2024, 482, 148885. [Google Scholar] [CrossRef]
  87. Xu, Y.; Liu, J.; Xie, M.; Jing, L.; Xu, H.; She, X.; Li, H.; Xie, J. Construction of Novel CNT/LaVO4 Nanostructures for Efficient Antibiotic Photodegradation. Chem. Eng. J. 2019, 357, 487–497. [Google Scholar] [CrossRef]
  88. Hammad, M.; Angel, S.; Al-kamal, A.K.; Asghar, A.; Kräenbring, M.A.; Amin, A.; Wiedemann, H.T.A.; Amin, A.S.; Vinayakumar, V.; Schmidt, T.C.; et al. Spray-Flame Synthesis of LaCo0.2Mn0.8O3 for Selective Peroxymonosulfate Activation into Singlet Oxygen towards Efficient Degradation of Carbamazepine. Process Saf. Environ. Prot. 2025, 194, 1347–1359. [Google Scholar] [CrossRef]
  89. Mafa, P.J.; Malefane, M.E.; Idris, A.O.; Mamba, B.B.; Liu, D.; Gui, J.; Kuvarega, A.T. Cobalt Oxide/Copper Bismuth Oxide/Samarium Vanadate (Co3O4/CuBi2O4/SmVO4) Dual Z-Scheme Heterostructured Photocatalyst with High Charge-Transfer Efficiency: Enhanced Carbamazepine Degradation under Visible Light Irradiation. J. Colloid. Interface Sci. 2021, 603, 666–684. [Google Scholar] [CrossRef] [PubMed]
  90. Pan, F.; Ji, H.; Du, P.; Huang, T.; Wang, C.; Liu, W. Insights into Catalytic Activation of Peroxymonosulfate for Carbamazepine Degradation by MnO2 Nanoparticles In-Situ Anchored Titanate Nanotubes: Mechanism, Ecotoxicity and DFT Study. J. Hazard Mater. 2021, 402, 123779. [Google Scholar] [CrossRef]
  91. Chen, G.; Wang, H.; Dong, W.; Huang, Y.; Zhao, Z.; Zeng, Y. Graphene Dispersed and Surface Plasmon Resonance-Enhanced Ag3PO4 (DSPR-Ag3PO4) for Visible Light Driven High-Rate Photodegradation of Carbamazepine. Chem. Eng. J. 2021, 405, 126850. [Google Scholar] [CrossRef]
  92. Zhou, Y.; Zhou, L.; Ni, C.; He, E.; Yu, L.; Li, X. 3D/2D MOF-Derived CoCeOx/g-C3N4 Z-Scheme Heterojunction for Visible Light Photocatalysis: Hydrogen Production and Degradation of Carbamazepine. J. Alloys Compd. 2022, 890, 161786. [Google Scholar] [CrossRef]
  93. Qudsieh, I.Y.; Ali, M.A.; Maafa, I.M. Effect of Water Matrix on Photocatalytic Degradation of Organic Pollutants in Water: A Literature Review. Rev. Chem. Eng. 2025, 41, 539–573. [Google Scholar] [CrossRef]
  94. Heredia Deba, S.A.; Wols, B.A.; Yntema, D.R.; Lammertink, R.G.H. Effects of the Water Matrix on the Degradation of Micropollutants by a Photocatalytic Ceramic Membrane. Membranes 2022, 12, 1004. [Google Scholar] [CrossRef]
  95. Pavel, M.; Anastasescu, C.; State, R.N.; Vasile, A.; Papa, F.; Balint, I. Photocatalytic Degradation of Organic and Inorganic Pollutants to Harmless End Products: Assessment of Practical Application Potential for Water and Air Cleaning. Catalysts 2023, 13, 380. [Google Scholar] [CrossRef]
  96. Gaber, M.M.; Samy, M.; Azam, A.; Shokry, H. Remediation of Paracetamol-Contaminated Water by Novel Burned Hookah Charcoal Residues via Persulfate Activation under Visible Light: Optimization, Mechanism, and Real Pharmaceutical Wastewater Application. J. Environ. Chem. Eng. 2024, 12, 114399. [Google Scholar] [CrossRef]
  97. Gaber, M.M.; Samy, M.; Shokry, H. Effective Degradation of Synthetic Micropollutants and Real Textile Wastewater via a Visible Light-Activated Persulfate System Using Novel Spinach Leaf-Derived Biochar. Environ. Sci. Pollut. Res. 2024, 31, 25163–25181. [Google Scholar] [CrossRef] [PubMed]
  98. Fouad, M.; Gar Alalm, M.; El-Etriby, H.K.; Boffito, D.C.; Ookawara, S.; Ohno, T.; Fujii, M. Visible-Light-Driven Photocatalytic Disinfection of Raw Surface Waters (300–5000 CFU/ML) Using Reusable Coated Ru/WO3/ZrO2. J. Hazard. Mater. 2021, 402, 123514. [Google Scholar] [CrossRef] [PubMed]
  99. Wen, J.; Zhou, L.; Tang, Q.; Xiao, X.; Sun, S. Photocatalytic Degradation of Organic Pollutants by Carbon Quantum Dots Functionalized G-C3N4: A Review. Ecotoxicol. Environ. Saf. 2023, 262, 115133. [Google Scholar] [CrossRef]
  100. Alalm, M.G.; Djellabi, R.; Meroni, D.; Pirola, C.; Bianchi, C.L.; Boffito, D.C. Toward Scaling-Up Photocatalytic Process for Multiphase Environmental Applications. Catalysts 2021, 11, 562. [Google Scholar] [CrossRef]
  101. Cheng, R.; Xia, J.C.; Shen, L.J.; Shen, Z.P.; Shi, L.; Zheng, X.; Zheng, J.Z. Effect of Humic Acid on Visible Light Photocatalytic Inactivation of Bacteriophage F2 with Electrospinning Cu-TiO2 Nanofibers: Insight into the Mechanisms. Environ. Sci. Pollut. Res. 2024, 31, 30212–30227. [Google Scholar] [CrossRef]
  102. Hu, M.; Chen, W.; Wang, J. Photocatalytic Degradation of Tetracycline by La-Fe Co-Doped SrTiO3/TiO2 Composites: Performance and Mechanism Study. Water 2024, 16, 210. [Google Scholar] [CrossRef]
  103. Feng, R.; Chen, L.; Li, W.; Cai, T.; Jiang, C. Activation of Persulfate with Natural Organic Acids (Ascorbic Acid/Catechin Hydrate) for Naproxen Degradation in Water and Soil: Mechanism, Pathway, and Toxicity Assessment. J. Hazard. Mater. 2023, 459, 132152. [Google Scholar] [CrossRef]
  104. Yang, H.; Joo, J.; Hong, E.; Park, S.J.; Lee, J.; Lee, C.G. Chicken Litter-Derived Catalyst for Persulfate Activation to Remove Acetaminophen: An Organic-Waste-to-Wealth Strategy. Chem. Eng. J. 2023, 471, 144368. [Google Scholar] [CrossRef]
  105. Tian, C.; Zhao, H.; Mei, J.; Yang, S. Cost-Efficient Graphitic Carbon Nitride as an Effective Photocatalyst for Antibiotic Degradation: An Insight into the Effects of Different Precursors and Coexisting Ions, and Photocatalytic Mechanism. Chem. Asian J. 2019, 14, 162–169. [Google Scholar] [CrossRef]
  106. Pete, K.Y.; Kabuba, J.; Otieno, B.; Ochieng, A. Modeling Adsorption and Photocatalytic Treatment of Recalcitrant Contaminant on Multi-Walled Carbon/TiO2 Nanocomposite. Environ. Sci. Pollut. Res. 2023, 30, 94154–94165. [Google Scholar] [CrossRef]
  107. Stavrinou, A.; Theodoropoulou, M.A.; Tsakiroglou, C.D. Synthesis of Titania/Activated Carbon Composites for the Synergistic Adsorption and Photocatalysis of Lindane in Aqueous Solutions. Environ. Sci. Pollut. Res. 2025, 32, 6468–6491. [Google Scholar] [CrossRef]
  108. Xu, W.; Wang, Q.; He, J.; Liu, F.; Yan, X.; Xu, Y. Visible Light-Assisted Periodate Activation Using Carbon Nitride for the Efficient Elimination of Acid Orange 7. Separations 2024, 11, 274. [Google Scholar] [CrossRef]
  109. Zhang, X.; Kamali, M.; Yu, X.; Costa, M.E.V.; Appels, L.; Cabooter, D.; Dewil, R. Kinetics and Mechanisms of the Carbamazepine Degradation in Aqueous Media Using Novel Iodate-Assisted Photochemical and Photocatalytic Systems. Sci. Total Environ. 2022, 825, 153871. [Google Scholar] [CrossRef]
  110. Elmitwalli, T.; Fouad, M.; Mossad, M.; Samy, M. Periodate Activation by Mulukhiyah Stalks and Potato Peels-Derived Biochars for the Efficient Degradation of Sulfamethazine. J. Environ. Chem. Eng. 2024, 12, 112101. [Google Scholar] [CrossRef]
  111. Mohamed, A.; Mahanna, H.; Samy, M. Synergistic Effects of Photocatalysis-Periodate Activation System for the Degradation of Emerging Pollutants Using GO/MgO Nanohybrid. J. Environ. Chem. Eng. 2024, 12, 112248. [Google Scholar] [CrossRef]
  112. Moradian, F.; Ramavandi, B.; Jaafarzadeh, N.; Kouhgardi, E. Activation of Periodate Using Ultrasonic Waves and UV Radiation for Landfill Leachate Treatment. Environ. Sci. Pollut. Res. 2022, 29, 90338–90350. [Google Scholar] [CrossRef] [PubMed]
  113. Ling, C.; Wu, S.; Han, J.; Dong, T.; Zhu, C.; Li, X.; Xu, L.; Zhang, Y.; Zhou, M.; Pan, Y. Sulfide-Modified Zero-Valent Iron Activated Periodate for Sulfadiazine Removal: Performance and Dominant Routine of Reactive Species Production. Water Res. 2022, 220, 118676. [Google Scholar] [CrossRef]
  114. Guo, D.; Yao, Y.; You, S.; Jin, L.; Lu, P.; Liu, Y. Ultrafast Degradation of Micropollutants in Water via Electro-Periodate Activation Catalyzed by Nanoconfined Fe2O3. Appl. Catal. B 2022, 309, 121289. [Google Scholar] [CrossRef]
  115. Monteagudo, J.M.; Durán, A.; San Martín, I.; Carrillo, P. Effect of Sodium Persulfate as Electron Acceptor on Antipyrine Degradation by Solar TiO2 or TiO2/RGO Photocatalysis. Chem. Eng. J. 2019, 364, 257–268. [Google Scholar] [CrossRef]
  116. Su, S.; Liu, Y.; He, W.; Tang, X.; Jin, W.; Zhao, Y. A Novel Graphene Oxide-Carbon Nanotubes Anchored α-FeOOH Hybrid Activated Persulfate System for Enhanced Degradation of Orange II. J. Environ. Sci. 2019, 83, 73–84. [Google Scholar] [CrossRef]
  117. Zhang, Y.; Zhang, C.; Zhou, Z.; Wu, Y.; Xing, S. Degradation of Ciprofloxacin by Persulfate Activation with CuO Supported on Mg Al Layered Double Hydroxide. J. Environ. Chem. Eng. 2021, 9, 106178. [Google Scholar] [CrossRef]
  118. Niu, L.; Zhang, G.; Xian, G.; Ren, Z.; Wei, T.; Li, Q.; Zhang, Y.; Zou, Z. Tetracycline Degradation by Persulfate Activated with Magnetic γ-Fe2O3/CeO2 Catalyst: Performance, Activation Mechanism and Degradation Pathway. Sep. Purif. Technol. 2021, 259, 118156. [Google Scholar] [CrossRef]
  119. Zhang, H.; Mei, Y.; Zhu, F.; Yu, F.; Komarneni, S.; Ma, J. Efficient Activation of Persulfate by C@Fe3O4 in Visible-Light for Tetracycline Degradation. Chemosphere 2022, 306, 135635. [Google Scholar] [CrossRef]
  120. Samy, M.; Mensah, K.; El-Fakharany, E.M.; Elkady, M.; Shokry, H. Green Valorization of End-of-Life Toner Powder to Iron Oxide-Nanographene Nanohybrid as a Recyclable Persulfate Activator for Degrading Emerging Micropollutants. Environ. Res. 2023, 223, 115460. [Google Scholar] [CrossRef]
  121. Zhao, X.; Wang, Y.; Liu, F.; Ye, X.; Wei, S.; Sun, Y.; He, J. Synergistic Photocatalytic Oxidation and Reductive Activation of Peroxymonosulfate by Bi-Based Heterojunction for Highly Efficient Organic Pollutant Degradation. Nanomaterials 2025, 15, 471. [Google Scholar] [CrossRef]
  122. Ali, N.; Khan, A.A.; Wakeel, M.; Khan, I.A.; Din, S.U.; Qaisrani, S.A.; Khan, A.M.; Hameed, M.U. Activation of Peroxymonosulfate by UV-254 Nm Radiation for the Degradation of Crystal Violet. Water 2022, 14, 3440. [Google Scholar] [CrossRef]
  123. Wang, G.; Hambly, A.C.; Zhao, D.; Wang, G.; Tang, K.; Andersen, H.R. Peroxymonosulfate Activation by Suspended Biogenic Manganese Oxides for Polishing Micropollutants in Wastewater Effluent. Sep. Purif. Technol. 2023, 306, 122501. [Google Scholar] [CrossRef]
  124. Bouzayani, B.; Lomba-Fernández, B.; Fdez-Sanromán, A.; Elaoud, S.C.; Sanromán, M.Á. Advancements in Copper-Based Catalysts for Efficient Generation of Reactive Oxygen Species from Peroxymonosulfate. Appl. Sci. 2024, 14, 8075. [Google Scholar] [CrossRef]
  125. Liu, T.; Li, C.X.; Chen, X.; Chen, Y.; Cui, K.; Wei, Q. Peroxymonosulfate Activation by Rice-Husk-Derived Biochar (RBC) for the Degradation of Sulfamethoxazole: The Key Role of Hydroxyl Groups. Int. J. Mol. Sci. 2024, 25, 11582. [Google Scholar] [CrossRef] [PubMed]
  126. Sun, H.; Peng, X.; Zhang, S.; Liu, S.; Xiong, Y.; Tian, S.; Fang, J. Activation of Peroxymonosulfate by Nitrogen-Functionalized Sludge Carbon for Efficient Degradation of Organic Pollutants in Water. Bioresour. Technol. 2017, 241, 244–251. [Google Scholar] [CrossRef]
  127. Cardoso, I.M.F.; Cardoso, R.M.F.; Esteves da Silva, J.C.G. Advanced Oxidation Processes Coupled with Nanomaterials for Water Treatment. Nanomaterials 2021, 11, 2045. [Google Scholar] [CrossRef]
  128. Mohammed, N.; Palaniandy, P.; Shaik, F.; Mewada, H.; Balakrishnan, D. Comparative Studies of RSM Box-Behnken and ANN-Anfis Fuzzy Statistical Analysis for Seawater Biodegradability Using TiO2 Photocatalyst. Chemosphere 2023, 314, 137665. [Google Scholar] [CrossRef]
  129. Yousefi, N.; Nabizadeh, R.; Nasseri, S.; Khoobi, M.; Nazmara, S.; Mahvi, A.H. Optimization of the Synthesis and Operational Parameters for NOM Removal with Response Surface Methodology during Nano-Composite Membrane Filtration. Water Sci. Technol. 2018, 77, 1558–1569. [Google Scholar] [CrossRef]
  130. Gaber, M.M.; Shokry, H.; Hassanin, A.H.; Awad, S.; Samy, M.; Elkady, M. Novel Palm Peat Lignocellulosic Adsorbent Derived from Agricultural Residues for Efficient Methylene Blue Dye Removal from Textile Wastewater. Appl. Water Sci. 2025, 15, 32. [Google Scholar] [CrossRef]
  131. Karimifard, S.; Alavi Moghaddam, M.R. Application of Response Surface Methodology in Physicochemical Removal of Dyes from Wastewater: A Critical Review. Sci. Total Environ. 2018, 640, 772–797. [Google Scholar] [CrossRef]
  132. Mirzaei, A.; Yerushalmi, L.; Chen, Z.; Haghighat, F.; Guo, J. Enhanced Photocatalytic Degradation of Sulfamethoxazole by Zinc Oxide Photocatalyst in the Presence of Fluoride Ions: Optimization of Parameters and Toxicological Evaluation. Water Res. 2018, 132, 241–251. [Google Scholar] [CrossRef]
  133. Diab, K.E.; Salama, E.; Hassan, H.S.; Abd El-moneim, A.; Elkady, M.F. Biocompatible MIP-202 Zr-MOF Tunable Sorbent for Cost-Effective Decontamination of Anionic and Cationic Pollutants from Waste Solutions. Sci. Rep. 2021, 11, 6619. [Google Scholar] [CrossRef] [PubMed]
  134. Naciri, Y.; Ahdour, A.; Benhsina, E.; Hamza, M.A.; Bouziani, A.; Hsini, A.; Bakiz, B.; Navío, J.A.; Ghazzal, M.N. Ba3(PO4)2 Photocatalyst for Efficient Photocatalytic Application. Glob. Chall. 2024, 8, 2300257. [Google Scholar] [CrossRef] [PubMed]
  135. Rahimi, B.; Jafari, N.; Abdolahnejad, A.; Farrokhzadeh, H.; Ebrahimi, A. Application of Efficient Photocatalytic Process Using a Novel BiVO/TiO2-NaY Zeolite Composite for Removal of Acid Orange 10 Dye in Aqueous Solutions: Modeling by Response Surface Methodology (RSM). J Environ Chem Eng 2019, 7, 103253. [Google Scholar] [CrossRef]
  136. Zhao, F.; Liu, Y.; Hammouda, S.B.; Doshi, B.; Guijarro, N.; Min, X.; Tang, C.J.; Sillanpää, M.; Sivula, K.; Wang, S. MIL-101(Fe)/g-C3N4 for Enhanced Visible-Light-Driven Photocatalysis toward Simultaneous Reduction of Cr(VI) and Oxidation of Bisphenol A in Aqueous Media. Appl. Catal. B. 2020, 272, 119033. [Google Scholar] [CrossRef]
  137. Gao, Y.; Xu, H.; Yin, K.; Li, Z.; Fan, J.; Wu, X.; Wu, Z. Modulating Charge Transfer for Photo-Driven N2O Reduction via Electronegativity Differences between Cu and Support. Appl. Catal. B Environ. Energy 2025, 377, 125528. [Google Scholar] [CrossRef]
  138. Goudjil, M.B.; Dali, H.; Zighmi, S.; Mahcene, Z.; Bencheikh, S.E. Photocatalytic Degradation of Methylene Blue Dye with Biosynthesized Hematite α-Fe2O3 Nanoparticles under UV-Irradiation. Desalination Water Treat 2024, 317, 100079. [Google Scholar] [CrossRef]
  139. Roškaric, M.; Žerjav, G.; Zavašnik, J.; Pintar, A. The Influence of Synthesis Conditions on the Visible-Light Triggered Photocatalytic Activity of g-C3N4/TiO2 Composites Used in AOPs. J. Environ. Chem. Eng. 2022, 10, 107656. [Google Scholar] [CrossRef]
  140. Ni, W.; Khan, A. Modified Metal-Organic Frameworks as Photocatalysts. Metal-Organic Frameworks for Chemical Reactions; Elsevier: Amsterdam, The Netherlands, 2021; pp. 231–270. [Google Scholar] [CrossRef]
  141. Ran, Z.; Fang, Y.; Sun, J.; Ma, C.; Li, S. Photocatalytic Oxidative Degradation of Carbamazepine by TiO2 Irradiated by UV Light Emitting Diode. Catalysts 2020, 10, 540. [Google Scholar] [CrossRef]
  142. APHA. AWWA and WEF Standard Methods for the Examination of Water and Wastewater; American Public Works Association: Kansas City, MO, USA, 2017; Available online: https://yabesh.ir/wp-content/uploads/2018/02/Standard-Methods-23rd-Perv.pdf (accessed on 17 September 2025).
  143. Song, L.; Zhu, B.; Gray, S.; Duke, M.; Muthukumaran, S. Performance of Hybrid Photocatalytic-Ceramic Membrane System for the Treatment of Secondary Effluent. Membranes 2017, 7, 20. [Google Scholar] [CrossRef] [PubMed]
  144. Theodorakopoulos, G.V.; Papageorgiou, S.K.; Katsaros, F.K.; Romanos, G.E.; Beazi-Katsioti, M. Investigation of MO Adsorption Kinetics and Photocatalytic Degradation Utilizing Hollow Fibers of Cu-CuO/TiO2 Nanocomposite. Materials 2024, 17, 4663. [Google Scholar] [CrossRef] [PubMed]
  145. Tsaviv, J.N.; Eneji, I.S.; Shato’Ato, R.; Ahemen, I.; Jubu, P.R.; Yusof, Y. Photodegradation, Kinetics and Non-Linear Error Functions of Methylene Blue Dye Using SrZrO3 Perovskite Photocatalyst. Heliyon 2024, 10, e34517. [Google Scholar] [CrossRef]
  146. Pesik, S.; Jobiliong, E.; Steven, E. Comparative Analysis of Photodegradation of Ibuprofen and Clotrimazole Water Pollutant Using UVC Rays in Presence and Absence of ZnO Photocatalyst. Environ. Sci. Proc. 2023, 25, 49. [Google Scholar] [CrossRef]
Figure 1. XRD patterns of the synthesized materials: (a) ZnO nanoparticles compared with the standard ZnO diffraction data (JCPDS No. 00-36-1451), (b) MIP-202(Zr) MOF, and (c) ZMIP nanocomposite.
Figure 1. XRD patterns of the synthesized materials: (a) ZnO nanoparticles compared with the standard ZnO diffraction data (JCPDS No. 00-36-1451), (b) MIP-202(Zr) MOF, and (c) ZMIP nanocomposite.
Catalysts 15 01017 g001aCatalysts 15 01017 g001b
Figure 2. FTIR spectra of the synthesized materials.
Figure 2. FTIR spectra of the synthesized materials.
Catalysts 15 01017 g002
Figure 3. (a,b) TEM images at magnifications of 20 Kx and 40 Kx, respectively; (c) HRTEM micrograph; (d) EDS pattern and (e) elemental mapping images of the green-synthesized ZnO nanoparticles.
Figure 3. (a,b) TEM images at magnifications of 20 Kx and 40 Kx, respectively; (c) HRTEM micrograph; (d) EDS pattern and (e) elemental mapping images of the green-synthesized ZnO nanoparticles.
Catalysts 15 01017 g003
Figure 4. (a,b) TEM images at magnifications of 20 Kx and 40 Kx, respectively; (c) HRTEM micrograph; (d) EDS pattern and (e) elemental mapping images of the synthesized MIP-202(Zr) bio-MOF.
Figure 4. (a,b) TEM images at magnifications of 20 Kx and 40 Kx, respectively; (c) HRTEM micrograph; (d) EDS pattern and (e) elemental mapping images of the synthesized MIP-202(Zr) bio-MOF.
Catalysts 15 01017 g004
Figure 5. (a,b) TEM images at magnifications of 20 Kx and 40 Kx, respectively; (c) HRTEM micrograph; (d) EDS pattern and (e) elemental mapping images of the ZMIP composite.
Figure 5. (a,b) TEM images at magnifications of 20 Kx and 40 Kx, respectively; (c) HRTEM micrograph; (d) EDS pattern and (e) elemental mapping images of the ZMIP composite.
Catalysts 15 01017 g005
Figure 6. Estimated Eg values of (a) green-synthesized ZnO nanoparticles, (b) MIP-202(Zr) bio-MOF, and (c) ZMIP nanocomposite (The red curves represent the experimental (αhν)2 versus hν plot, while the blue straight lines indicate the linear portion used to estimate the optical band gap energy).
Figure 6. Estimated Eg values of (a) green-synthesized ZnO nanoparticles, (b) MIP-202(Zr) bio-MOF, and (c) ZMIP nanocomposite (The red curves represent the experimental (αhν)2 versus hν plot, while the blue straight lines indicate the linear portion used to estimate the optical band gap energy).
Catalysts 15 01017 g006
Figure 7. CBZ degradation efficiency under different systems (Experimental conditions: [CBZ]o = 20 mg/L, [ZMIP]o = 1 g/L, pH = 7, and reaction time = 60 min). The black line represents CBZ degradation in the ZMIP/Light system under the optimum conditions (initial CBZ concentration = 15 mg/L, ZMIP dosage = 1.25 g/L, pH = 6, and reaction time = 90 min).
Figure 7. CBZ degradation efficiency under different systems (Experimental conditions: [CBZ]o = 20 mg/L, [ZMIP]o = 1 g/L, pH = 7, and reaction time = 60 min). The black line represents CBZ degradation in the ZMIP/Light system under the optimum conditions (initial CBZ concentration = 15 mg/L, ZMIP dosage = 1.25 g/L, pH = 6, and reaction time = 90 min).
Catalysts 15 01017 g007
Figure 8. Contour diagrams represent the effects of the operational parameters on CBZ removal in the ZMIP/light system: (a) reaction time vs. ZMIP dose, (b) reaction time vs. CBZ concentration, (c) reaction time vs. pH, (d) pH vs. CBZ concentration, (e) pH vs. ZMIP dose, and (f) CBZ concentration vs. ZMIP dose.
Figure 8. Contour diagrams represent the effects of the operational parameters on CBZ removal in the ZMIP/light system: (a) reaction time vs. ZMIP dose, (b) reaction time vs. CBZ concentration, (c) reaction time vs. pH, (d) pH vs. CBZ concentration, (e) pH vs. ZMIP dose, and (f) CBZ concentration vs. ZMIP dose.
Catalysts 15 01017 g008
Figure 9. Kinetic Study of CBZ photodegradation in the ZMIP/light system at varying initial CBZ concentrations: (a) pseudo-first-order and (b) pseudo-second-order. Conditions: pH = 7, catalyst loading = 1 g/L, and reaction time = 60 min.
Figure 9. Kinetic Study of CBZ photodegradation in the ZMIP/light system at varying initial CBZ concentrations: (a) pseudo-first-order and (b) pseudo-second-order. Conditions: pH = 7, catalyst loading = 1 g/L, and reaction time = 60 min.
Catalysts 15 01017 g009
Figure 10. (a) Degradation efficiency of CBZ in the ZMIP/light system over five consecutive cycles; (b) TOC mineralization efficiency of CBZ in the ZMIP/light system under the optimum conditions.
Figure 10. (a) Degradation efficiency of CBZ in the ZMIP/light system over five consecutive cycles; (b) TOC mineralization efficiency of CBZ in the ZMIP/light system under the optimum conditions.
Catalysts 15 01017 g010
Figure 11. Proposed charge transfer and ROS generation mechanisms for CBZ degradation in the ZMIP/light system.
Figure 11. Proposed charge transfer and ROS generation mechanisms for CBZ degradation in the ZMIP/light system.
Catalysts 15 01017 g011
Figure 12. (a) Impact of the addition of different quenchers (1 mM) on CBZ removal ratios in the ZMIP/light system under the optimum conditions; (b) photodegradation efficiency of various pharmaceutical pollutants in the ZMIP/light system.
Figure 12. (a) Impact of the addition of different quenchers (1 mM) on CBZ removal ratios in the ZMIP/light system under the optimum conditions; (b) photodegradation efficiency of various pharmaceutical pollutants in the ZMIP/light system.
Catalysts 15 01017 g012
Figure 13. Degradation products and proposed degradation pathways of CBZ using the ZMIP photocatalyst under visible light irradiation.
Figure 13. Degradation products and proposed degradation pathways of CBZ using the ZMIP photocatalyst under visible light irradiation.
Catalysts 15 01017 g013
Figure 14. Influence of (a) different aqueous media and (b) varying concentrations of co-existing humic acid on CBZ degradation efficiency in the ZMIP/light system. Conditions: [CBZ]o = 15 mg/L, reaction time = 90 min, [photocatalyst]o = 1.25 g/L, pH = 6).
Figure 14. Influence of (a) different aqueous media and (b) varying concentrations of co-existing humic acid on CBZ degradation efficiency in the ZMIP/light system. Conditions: [CBZ]o = 15 mg/L, reaction time = 90 min, [photocatalyst]o = 1.25 g/L, pH = 6).
Catalysts 15 01017 g014
Figure 15. Impact of (a) different co-existing inorganic anions (10 mM) and (b) different co-existing inorganic cations (2 mM) on CBZ removal efficiency within the ZMIP/light photocatalytic system. Conditions: [CBZ]o = 15 mg/L, reaction time = 90 min, [ZMIP]o = 1.25 g/L, pH = 6).
Figure 15. Impact of (a) different co-existing inorganic anions (10 mM) and (b) different co-existing inorganic cations (2 mM) on CBZ removal efficiency within the ZMIP/light photocatalytic system. Conditions: [CBZ]o = 15 mg/L, reaction time = 90 min, [ZMIP]o = 1.25 g/L, pH = 6).
Catalysts 15 01017 g015
Figure 16. (a) TOC removal efficiency of untreated pharmaceutical wastewater over five consecutive cycles in the ZMIP/light system, and (b) impact of the addition of various inorganic oxidants (3 mM) on CBZ degradation efficiency in the proposed photocatalytic system.
Figure 16. (a) TOC removal efficiency of untreated pharmaceutical wastewater over five consecutive cycles in the ZMIP/light system, and (b) impact of the addition of various inorganic oxidants (3 mM) on CBZ degradation efficiency in the proposed photocatalytic system.
Catalysts 15 01017 g016
Table 1. Experimental design parameters along with the corresponding actual and predicted CBZ removal efficiencies.
Table 1. Experimental design parameters along with the corresponding actual and predicted CBZ removal efficiencies.
RunCoded Values of ParametersActual Values of ParametersCBZ Removal (%)
DFGHD FGHActualPredicted
1−10−1030715186.8782.43
2−1−10030320163.0166.84
3000060720176.5876.49
4−1100301120149.2351.49
501−10601115165.7467.12
6−1001307201.2583.1784.08
71100901120172.6571.18
801016011201.2570.0068.07
9000060720176.5476.49
100110601125149.3448.18
11010−16011200.7548.9049.74
1210−1090715193.1293.44
130−1−1060315176.7378.64
141−10090320179.3279.43
150−101603201.2585.7781.83
16100−1907200.7580.0879.93
17000060720176.5976.49
180011607251.2577.0780.53
19000060720176.3276.49
201001907201.2596.2595.93
2100−11607151.2590.1892.10
220−10−1603200.7560.7559.58
23−101030725162.0558.63
240−11060325160.8860.25
25−100−1307200.7558.4359.50
26001−1607250.7552.6853.14
27101090725178.5679.90
2800−1−1607150.7580.0078.91
29000060720176.4176.49
Table 2. Derived optimum levels of the independent variables and CBZ removal efficiency under the optimum conditions.
Table 2. Derived optimum levels of the independent variables and CBZ removal efficiency under the optimum conditions.
ParameterOptimum Value
Reaction time (min)90
pH6
Initial CBZ concentration (mg/L)15
Photocatalyst dose (g/L)1.25
CBZ removal efficiency (Laboratory)99.37%
CBZ removal efficiency (Model)98.84%
Table 3. Photocatalytic efficiencies of diverse metal oxide–MOF hybrid photocatalysts for the degradation of various pharmaceutical contaminants.
Table 3. Photocatalytic efficiencies of diverse metal oxide–MOF hybrid photocatalysts for the degradation of various pharmaceutical contaminants.
PhotocatalystLight SourcePollutantOperating ConditionsRemoval Ratio (%)Reference
Bi4O7/MIL-68(In)-NH2Visible light: Xenon lamp (300 W)CBZ[Pollutant]o = 50 mg/L,
[Catalyst]o = 1 g/L, and reaction time = 120 min.
92.7[45]
TiO2/MIL-101(Cr)UV-A LED (20 W)CBZ[Pollutant]o = 12 mg/L,
[Catalyst]o = 2 g/L, and reaction time = 60 min.
99[46]
ZnO/ZIF-9Ultraviolet lightTetracycline[Pollutant]o = 20 mg/L,
[Catalyst]o = 0.25 g/L, and reaction time = 60 min.
87.7[47]
ZnO/UiO-66/NH2Visible light: Xenon lamp (50 W, 0.01–105 HZ)Tetracycline[Pollutant]o = 20 mg/L,
[Catalyst]o = 0.1 g/L, and reaction time = 90 min.
61[48]
ZnO/MIL-53(Al)Visible light: Metal-halide lamp (Venture, 400 W) Trimethoprim[Pollutant]o = 10 mg/L, pH = 7, flowrate = 5 mL/min, and reaction time = 240 min.93.5[49]
ZnO-CuO/TMU-5Visible light: Metal halide lamp (400 W, λ = 510 nm)Tetracycline[Pollutant]o = 30 mg/L,
[Catalyst]o = 1 g/L, and reaction time = 180 min.
62[50]
GO-WO3/UiO-66Visible light: Mercury lamp (400 W)Tetracycline[Pollutant]o = 20 mg/L,
[Catalyst]o = 1.67 g/L,
pH = 7, and reaction time = 70 min.
84[51]
TiO2/MIL-101(Cr)
Visible light: Xenon lamp (300 W, λ > 400 nm)Tetracycline[Pollutant]o = 10 mg/L,
[Catalyst]o = 0.2 g/L, and reaction time = 45 min.
94.85[52]
TiO2/UiO-66-NH2Visible light: Xenon lamp (350 W)Tetracycline[Pollutant]o = 20 mg/L,
[Catalyst]o = 0.12 mg TCN/mg catalyst, and reaction time = 60 min.
75[53]
Bi2WO6/MIL-88B(Fe)Visible light: Xenon lamp (500 W)Tetracycline[Pollutant]o = 10 mg/L,
[Catalyst]o = 50 mg/L, and reaction time = 90 min.
96.4[54]
Bi2WO6/MIL-125 (Ti)-NH2Visible light: Xenon lamp (300 W)Tetracycline[Pollutant]o = 20 mg/L,
[Catalyst]o = 0.4 g/L, and reaction time = 120 min.
69.19[55]
ZnO/MIP-202(Zr)Visible light: Metal halide lamp (400 W, λ = 510 nm)CBZ[Pollutant]o = 15 mg/L,
[Catalyst]o = 1.25 g/L,
pH = 6, and reaction time = 90 min.
99.37This study
Table 4. ANOVA of the quadratic model for CBZ degradation response.
Table 4. ANOVA of the quadratic model for CBZ degradation response.
SourceDFSum of SquaresMean SquareF-Valuep-Value
Model144760.41340.0346.180
Linear43481.67870.42118.20
D-Reaction time (min)1787.64787.64106.960
F-pH1415.36415.3656.410
G-CBZ concentration (mg/L)11046.451046.45142.110
H-Catalyst dose (g/L)11232.211232.21167.330
Square41167.02291.7639.620
D (min) × D (min)154.5354.537.40.017
F × F1958.05958.05130.10
G (mg/L) × G (mg/L)14.064.060.550.47
H (g/L) × H (g/L)11.411.410.190.668
2-Way Interaction6111.7118.622.530.072
D (min) × F112.6412.641.720.211
D (min) × G (mg/L)126.3226.323.570.08
D (min) × H (g/L)118.3618.362.490.137
F × G (mg/L)10.080.080.010.921
F × H (g/L)13.843.840.520.482
G (mg/L) × H (g/L)150.4850.486.860.02
Error14103.097.36
Lack-of-Fit10103.0410.3737.560
Pure Error40.060.01
Total284863.5
R2 = 97.88%, R2 (adjusted) = 95.76%.
Table 5. Factors and design levels for the photocatalytic degradation of CBZ in the ZMIP/light system.
Table 5. Factors and design levels for the photocatalytic degradation of CBZ in the ZMIP/light system.
Independent ParameterCodeUnitLevels
−101
Reaction timeDmin306090
pHF3711
Initial CBZ concentrationGmg/L152025
Photocatalyst doseHg/L0.7511.25
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

Gaber, M.M.; Toghan, A.; Eldesoky, A.M.; Al-Hussain, S.A.; Masoud, E.M.; Shokry, H.; Samy, M.; Elkady, M. Sustainable Photocatalytic Treatment of Real Pharmaceutical Wastewater Using a Novel ZnO/MIP-202(Zr) Bio-MOF Hybrid Synthesized via a Green Approach. Catalysts 2025, 15, 1017. https://doi.org/10.3390/catal15111017

AMA Style

Gaber MM, Toghan A, Eldesoky AM, Al-Hussain SA, Masoud EM, Shokry H, Samy M, Elkady M. Sustainable Photocatalytic Treatment of Real Pharmaceutical Wastewater Using a Novel ZnO/MIP-202(Zr) Bio-MOF Hybrid Synthesized via a Green Approach. Catalysts. 2025; 15(11):1017. https://doi.org/10.3390/catal15111017

Chicago/Turabian Style

Gaber, Mohamed Mohamed, Arafat Toghan, Ahmed M. Eldesoky, Sami A. Al-Hussain, Emad M. Masoud, Hassan Shokry, Mahmoud Samy, and Marwa Elkady. 2025. "Sustainable Photocatalytic Treatment of Real Pharmaceutical Wastewater Using a Novel ZnO/MIP-202(Zr) Bio-MOF Hybrid Synthesized via a Green Approach" Catalysts 15, no. 11: 1017. https://doi.org/10.3390/catal15111017

APA Style

Gaber, M. M., Toghan, A., Eldesoky, A. M., Al-Hussain, S. A., Masoud, E. M., Shokry, H., Samy, M., & Elkady, M. (2025). Sustainable Photocatalytic Treatment of Real Pharmaceutical Wastewater Using a Novel ZnO/MIP-202(Zr) Bio-MOF Hybrid Synthesized via a Green Approach. Catalysts, 15(11), 1017. https://doi.org/10.3390/catal15111017

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop