Next Article in Journal
The Impact of Sm Promoter on the Catalytic Performance of Ni/Al2O3-SiO2 in Methane Partial Oxidation for Enhanced H2 Production
Previous Article in Journal
High-Performance Catalytic Oxygen Evolution with Nanocellulose-Derived Biocarbon and Fe/Zeolite/Carbon Nanotubes
Previous Article in Special Issue
Morphological Regulation of Bi5O7I for Enhanced Efficiency of Rhodamine B Degradation Under Visible-Light
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Tunable Optical Bandgap and Enhanced Visible Light Photocatalytic Activity of ZnFe2O3-Doped ZIF-8 Composites for Sustainable Environmental Remediation

1
Department of Physics, College of Science, Princess Nourah bint Abdulrahman University, P.O. Box 84428, Riyadh 11671, Saudi Arabia
2
Advanced Materials Research Laboratory, Department of Physics, Faculty of Science, University of Tabuk, Tabuk 71491, Saudi Arabia
*
Authors to whom correspondence should be addressed.
Catalysts 2025, 15(8), 720; https://doi.org/10.3390/catal15080720
Submission received: 19 June 2025 / Revised: 26 July 2025 / Accepted: 26 July 2025 / Published: 29 July 2025
(This article belongs to the Special Issue Catalysis Accelerating Energy and Environmental Sustainability)

Abstract

Metal–organic frameworks (MOFs), particularly ZIF-8, have emerged as promising materials due to their high porosity, tunability, and chemical stability. In this study, we report the synthesis of ZnFe2O3-doped ZIF-8 composites with 10 wt% loading via a solvothermal method to enhance their optical and photocatalytic performance. Structural analyses confirmed the successful incorporation of ZnFe2O3 without disrupting the ZIF-8 framework. Optical studies revealed enhanced absorption in the visible range, a narrowed bandgap (4.26 eV vs. 4.37 eV for pristine ZIF-8), and an increased extinction coefficient, indicating superior light-harvesting potential. The photocatalytic activity was evaluated by methylene blue (MB) degradation under visible light, where the 10 wt% ZnFe2O3-ZIF-8 composite achieved 90% degradation efficiency, outperforming pristine ZIF-8 (67.8%). The catalyst also demonstrated excellent recyclability over five cycles and a proposed degradation mechanism involving ·OH and ·O2 radical formation. These findings demonstrate the potential of highly doped ZnFe2O3@ZIF-8 composites for environmental remediation and photonic applications.

Graphical Abstract

1. Introduction

Metal–organic frameworks (MOFs) have garnered significant attention recently due to their remarkable structural properties, tunable porosity, and extensive surface areas [1]. These properties make MOFs suitable for various applications in gas storage [2], separation [3], catalysis [4], and sensing [5,6]. Among the diverse types of MOFs, zeolitic imidazolate frameworks (ZIFs) stand out due to their unique high stability and versatility [7]. The prototypical member of this family is ZIF-8, which is particularly notable for its excellent thermal and chemical stability, as well as its intrinsic luminescent properties [8]. These characteristics position ZIF-8 as a promising candidate for innovative applications in optoelectronics [9,10], photonics [11], energy storage [12], and wastewater pollutant removal [13,14,15,16].
Despite its advantageous properties, pristine ZIF-8 often exhibits limited photocatalytic activity due to rapid electron–hole recombination and insufficient light absorption in the visible spectrum. Recent studies have explored the incorporation of metal oxides or nanoparticles into metal–organic frameworks (MOFs) to address these limitations and enhance their optical and photocatalytic properties. Zinc ferrite (ZnFe2O3), a mixed-metal oxide, has attracted considerable interest due to its high thermal stability, superior light absorption in the visible region, and ability to facilitate charge separation. Introducing ZnFe2O3 into MOFs can further enhance their properties, leading to composite materials with superior functionality [17]. ZnFe2O3 has attracted considerable interest due to its favorable physicochemical attributes, including high stability, effective light absorption, and magnetic properties [18]. The incorporation of ZnFe2O3 into ZIF-8 matrices not only fortifies the structural integrity but also synergistically augments the optical and electrochemical performance of the composites. This interplay between organic and inorganic components can unlock novel energy conversion and storage applications, particularly in supercapacitors and photonic devices [19].
Recent literature reveals a growing body of research focused on the optical properties and laser-limiting capabilities of various metal–organic frameworks (MOFs) and their composites. Studies indicate that incorporating metal oxides can significantly modify the absorption spectrum and enhance nonlinear optical properties, providing pathways for developing advanced materials that can mitigate laser damage and improve performance in photonic applications [20]. Furthermore, investigations into the energy gap of these composite materials have shown that doping can lead to a shift in electronic transitions, promoting enhanced performance in light-harvesting and storage applications [21].
Previous studies, including our recent work on 5 wt% ZnFe2O3-doped ZIF-8, have demonstrated improved electrochemical and supercapacitor behavior of ZIF-8 composites [22]. In this work, we advance the investigation by exploring a higher doping level (10 wt%) of ZnFe2O3 in ZIF-8, specifically targeting optical tuning and enhanced photocatalytic degradation of methylene blue (MB) under visible light. The emphasis shifts from electrochemical storage to photocatalysis and photonic applications, with a detailed assessment of band structure, extinction coefficient, and degradation kinetics. This study offers new insights into the structure–property relationships at elevated dopant levels, positioning ZnFe2O3@ZIF-8 as a promising platform for environmental remediation technologies.

2. Results and Discussion

2.1. Structural Properties

XRD is a crucial tool for examining the internal structure of novel composites. To look more closely at how the structure changes with more ZnFe2O3, Figure 1 shows the XRD patterns for ZIF-8, ZnFe2O3@ZIF-8 (5 wt%) [22], and ZnFe2O3@ZIF-8 (10 wt%). All samples display reflections linked to the ZIF-8 framework; however, the diffraction patterns of the doped composites, especially the 10 wt% sample, show notable deviations from the standard ZIF-8 patterns documented in the literature [23,24] and PDF card No. 00-062-1030. Most notably, the sharp and intense peaks at 2θ ≈ 31.7° and 36.9° are not present in standard ZIF-8 but are consistent with the crystalline phases of ZnFe2O3 or associated oxides, as indexed in PDF card No. 00-022-1012. This confirms the formation of secondary oxide phases and suggests partial structural transformation or surface crystallite formation at higher dopant concentrations. While some characteristic ZIF-8 peaks remain (e.g., at ~7.3° and 12.8°), the overall pattern indicates that the introduction of ZnFe2O3 at elevated levels leads to the coexistence of multiple crystalline phases and partial distortion of the ZIF-8 lattice.
However, the 10 wt% sample shows broader and weaker peaks, especially at the (002) and (112) spots, and new peaks related to ZnFe2O3 appear around 35.5° [25]. Compared to the 5 wt% sample, the difference suggests a higher degree of lattice distortion and secondary phase formation at higher dopant loading. These structural changes indicate that more ZnFe2O3 nanoparticles are being added, which correlates with the improved photocatalytic performance observed in the 10 wt% sample. In our previous work [22], we reported the structural behavior of ZIF-8 doped with 5 wt% ZnFe2O3, where the composite maintained good crystallinity with moderate peak broadening and minor phase impurity. In the current study, a higher ZnFe2O3 loading was used, resulting in further suppression of the ZIF-8 peak intensities and more pronounced secondary oxide peaks. This indicates increased structural distortion and partial framework substitution, possibly due to the excess oxide interacting with the coordination environment of zinc centers. The broader diffraction peaks also suggest reduced crystallite size or increased lattice strain, which are common outcomes of metal oxide doping in MOFs at higher concentrations. Overall, the XRD results confirm that while the ZIF-8 framework remains intact, higher ZnFe2O3 doping levels induce additional phase signatures and structural perturbations, effects that were much less pronounced at the lower 5 wt% level [22]. These changes may enhance catalytic surface activity and facilitate electron transfer by introducing more defect sites and heterojunction interfaces, which are beneficial for photocatalytic applications.
The thermal stability and decomposition behavior of the ZIF-8@ZnFe2O3 composite were investigated using thermogravimetric analysis (TGA) and differential thermogravimetric analysis (DTA). Figure 2 presents the simultaneous TGA (black curve) and DTA (blue curve) results for the ZnFe2O3-doped ZIF-8 composite. The TGA curve exhibits a three-step weight loss profile: an initial minor mass loss (~5%) below 150 °C is attributed to the evaporation of physically adsorbed water and residual solvents, as indicated by a shallow endothermic peak in the DTA curve. A more pronounced weight loss (~20%) between 250 and 350 °C corresponds to the partial decomposition of the organic 2-methylimidazole linker, accompanied by a broad exothermic peak, suggesting catalytic degradation induced by ZnFe2O3 nanoparticles. At temperatures above 500 °C, a significant exothermic peak and a sharp mass loss (~35–40%) are observed, marking the collapse of the ZIF-8 framework and the transformation of the hybrid structure into stable metal oxide residues, such as ZnO, Fe2O3, or ZnFe2O4 spinel. The final residual mass beyond 600 °C confirms the presence of thermally stable inorganic components. Compared to pristine ZIF-8, the composite exhibits enhanced thermal stability, likely due to interfacial interactions between the ZIF-8 framework and the ZnFe2O3 dopant. These results are in strong agreement with previous studies on similar MOF-based hybrids [22,26], reinforcing the suitability of the ZnFe2O3@ZIF-8 composite for applications involving thermal and photocatalytic processes.
On the other hand, the DTA thermogram exhibits distinct endothermic and exothermic transitions, providing insight into the thermal stability and decomposition behavior of the composite. A small endothermic peak observed around 100–150 °C corresponds to the removal of physically adsorbed moisture and residual solvents (e.g., DMF or methanol) retained during the synthesis process. This feature is typical for metal–organic frameworks and confirms the presence of weakly bound surface molecules [27,28]. In the intermediate temperature range (250–350 °C), a broad exothermic peak is apparent, which is associated with the partial decomposition of the organic 2-methylimidazole linker. The shift in this peak relative to pristine ZIF-8 suggests a catalytic effect introduced by ZnFe2O3, which facilitates the thermal degradation process. This behavior indicates the successful incorporation of ZnFe2O3 nanoparticles into the framework and their interaction with the host matrix. A more prominent exothermic peak appears at 510 °C, which is attributed to the complete collapse of the ZIF-8 framework and the formation of thermally stable metal oxide residues. The evolution of this peak at a slightly elevated temperature compared to pure ZIF-8 reflects an enhancement in thermal stability due to the presence of ZnFe2O3, potentially leading to the formation of spine-type ZnFe2O3 structures [27]. In our previous study [22], ZnFe2O3 was incorporated into ZIF-8 at a doping ratio of 5 wt%, demonstrating noticeable improvements in thermal behavior.
In contrast, the current study employs a higher doping level, which further enhances the thermal stability of the composite. The observed delay in decomposition temperature and slower mass loss rate confirms that increasing the ZnFe2O3 content strengthens the composite’s resistance to thermal degradation. This enhancement is likely due to intensified interactions between the metal oxide and the ZIF-8 matrix, which act as thermal buffers and stabilize the structure. These results align with previous reports [22,29] and reinforce the conclusion that the ZnFe2O3@ZIF-8 system, particularly at higher doping levels, is highly suitable for thermally demanding applications, such as photocatalysis under visible light.
The SEM image of pristine ZIF-8 (Figure 3a) reveals a typical morphology of rhombic dodecahedra, which is consistent with the reported crystal structure of ZIF-8 [30]. The surface of the ZIF-8 crystal is smooth and free of any impurities, indicating high crystallinity. The particle size distribution is uniform, with most particles falling in the range of 15 μm. In contrast, the SEM image of ZIF-8@ZnFe2O3 (Figure 3b) shows a significant change in the material’s morphology and surface characteristics. The introduction of ZnFe2O3 has led to the formation of a rough and porous surface, indicating the successful deposition of ZnFe2O3 on the ZIF-8 surface.
In contrast, the SEM image of ZIF-8@ZnFe2O3 (10 wt%) shows a noticeable change in morphology. The surface becomes rougher and more porous, indicating successful integration of ZnFe2O3 into the ZIF-8 structure. The particle size distribution becomes more heterogeneous, with some degree of aggregation observed. Notably, small spherical nanoparticles (average diameter ~43 nm), corresponding to ZnFe2O3, are visible and uniformly distributed on the ZIF-8 crystal surfaces, further supporting successful doping. These features confirm successful incorporation and suggest the creation of porous, high-surface-area sites that are ideal for photocatalysis.
Functional groups stacking on molecules have distinctive infrared spectra, and this property has made infrared spectroscopy the gold standard for physical investigations into molecular structure and functional group identification [31]. The Fourier transform infrared (FTIR) spectra of ZIF-8 and ZnFe2O3@ZIF-8 are presented in Figure 4. The FTIR spectrum confirms the successful synthesis and doping of ZIF-8 with ZnFe2O3. The key peaks at approximately 1580 cm−1 and 1140 cm−1 are related to the vibrations of the imidazole ring, indicating that the ZIF-8 structure remains intact after doping. Additionally, the peak at approximately 420 cm−1, attributed to Zn-N stretching, remains prominent in both pristine and doped samples, indicating that the primary framework is retained. Although we expected to see Fe-O or Zn-O stretching in the fingerprint region between 500 and 600 cm−1, the current spectra do not show any clear new peaks in that area. The broadband near ~3400 cm−1 observed in both spectra is attributed to O-H stretching vibrations of surface hydroxyl groups or physisorbed water molecules. These -OH groups may participate in photocatalytic processes by forming ·OH radicals, as previously reported [32]. The slight shift in some peaks (e.g., imidazole ring vibrations) suggests interactions between the ZnFe2O3 dopant and the ZIF-8 framework, potentially due to coordination or hydrogen bonding. The FTIR results demonstrate the successful doping of ZIF-8 with ZnFe2O3 while retaining the MOF’s essential framework structure. Incorporating ZnFe2O3 introduces metal–oxide-specific vibrations without disrupting the primary framework, thereby enhancing the composite’s potential for photocatalytic and energy storage applications.

2.2. Optical Properties

The optical transmission and reflection spectra of ZIF-8 and ZIF-8@ZnFe2O3 films are presented in Figure 5. The ZIF-8 film exhibits a smooth decline in transmission with increasing wavelength, indicative of its broad optical absorption. In contrast, the ZIF-8@ZnFe2O3 composite exhibits a more complex transmission profile, featuring distinct modulations and dips that signify the introduction of additional absorption bands resulting from the doping of ZnFe2O3. These features highlight the enhanced optical activity of the composite, which is attributed to the synergistic interaction between the ZIF-8 framework and ZnFe2O3 nanoparticles. The reflection spectra reveal low reflectivity for both films, consistent with the inherent properties of ZIF-8. However, the ZIF-8@ ZnFe2O3 composite demonstrates slightly increased reflectivity in specific regions due to surface morphology alterations or refractive index changes arising from the integration of ZnFe2O3. ZnFe2O3 is known for its strong visible-light absorption and charge-transfer properties, making it an excellent candidate for enhancing the photocatalytic efficiency of hybrid systems. The observed dips in the transmission spectrum indicate the formation of new energy states in ZnFe2O3, which may facilitate light harvesting and improve charge carrier separation within the composite. The results demonstrate that integrating ZnFe2O3 into the ZIF-8 framework effectively tunes its optical properties, making it a promising material for applications in photocatalysis and energy storage technologies.
The extinction coefficient (k) measures how much light is absorbed or scattered by a material at a specific wavelength. The absorption coefficient (α) and the extinction coefficient (k) can be determined by [33]
α = 1 x ln 1 R 2 2 T + R 2 + 1 R 2 4 T 2
k = λ × α/4π.
where x is the sample thickness, and R and T are the measured reflectance and transmittance, respectively.
The extinction coefficient spectra of ZIF-8 and ZIF-8@ZnFe2O3 films are shown in Figure 6. The extinction coefficient provides critical insights into the light absorption capability of the material and energy dissipation processes. For the ZIF-8 film, the extinction coefficient exhibits one prominent peak at shorter wavelengths, corresponding to electronic transitions within the ZIF-8 framework. This peak is followed by a gradual decline at higher wavelengths, indicating reduced optical absorption in the visible and near-infrared regions. In the case of the ZIF-8@ZnFe2O3 composite, the spectra show enhanced extinction in the shorter wavelength region, with a broader and more intense peak compared to pristine ZIF-8. This enhancement suggests that ZnFe2O3 doping introduces additional electronic states or transitions due to the formation of heterojunctions between the ZIF-8 framework and ZnFe2O3 nanoparticles. The extended tail in the visible region for ZIF-8@ZnFe2O3 further highlights its improved light-harvesting capabilities, making it a promising material for photocatalytic applications. The significant increase in k for ZIF-8@ZnFe2O3 suggests enhanced interfacial interaction between the two components, facilitating the absorption of high-energy photons and potentially improving photocatalytic performance. This behavior confirms that integrating ZnFe2O3 into the ZIF-8 framework is an effective strategy for tailoring its optical properties in advanced energy and photocatalytic applications.
The indirect energy bandgap of ZIF-8 and ZIF-8@ZnFe2O3 films was determined using Tauc plots, as shown in Figure 7. The Tauc relation was used to calculate the bandgap energy, which is given by [34]
α h υ 1 2 = Q   h υ E g
where Q is a constant. A bandgap of approximately 4.37 eV was found for ZIF-8. This value aligns with previous reports on ZIF-8 materials [35], confirming its wide bandgap semiconducting nature. In comparison, the ZIF-8@ZnFe2O3 composite exhibits a reduced bandgap of 4.26 eV, indicating the influence of ZnFe2O3 doping. The reduced Eg can be attributed to the introduction of mid-gap states and improved charge carrier transfer at the ZIF-8/ZnFe2O3 interface. This behavior aligns with studies on metal oxide-doped MOFs, where bandgap narrowing enhances light absorption and photocatalytic activity [36]. The narrowed bandgap of the composite is advantageous for photocatalysis under visible light, as it extends the absorption range of the material and promotes efficient charge separation. These findings further demonstrate the potential of ZIF-8@ ZnFe2O3 as a tailored material for advanced photocatalytic and energy applications.
The refractive index, n, can be calculated using [33]
n = 1 + R 1 R + 4 R 1 R 2 k 2
The refractive index spectra of ZIF-8 and ZnFe2O3-doped ZIF-8 are presented in Figure 8. Both samples exhibit a prominent peak at shorter wavelengths, which decreases smoothly as the wavelength increases. For pristine ZIF-8, the refractive index exhibits a sharp peak and then gradually diminishes, demonstrating typical dielectric behavior. In contrast, the ZnFe2O3-doped sample shows consistently higher refractive index values across the entire wavelength range.
The enhanced refractive index in ZIF-8@ZnFe2O3 can be attributed to the incorporation of ZnFe2O3 nanoparticles. These nanoparticles increase the material’s density and polarizability, thereby improving its optical response. The sharp peak observed at shorter wavelengths corresponds to electronic transitions and resonance effects, indicating enhanced light–matter interaction due to doping. Furthermore, the broader optical response observed in the doped sample suggests enhanced light-harvesting capabilities, which are beneficial for photocatalytic applications. The observed enhancement in the optical response of ZIF-8@ZnFe2O3, compared to pristine ZIF-8, highlights the synergistic effect of combining ZnFe2O3 with the porous ZIF-8 framework. This improved optical performance underscores the potential of such composites in advanced photocatalytic and energy storage applications.
The dielectric constant (ε1 = n2k2) and dielectric loss (ε2 = 2 nk) spectra for ZIF-8 and ZnFe2O3-doped ZIF-8 are depicted in Figure 9. Both ε1 and ε2 exhibit distinct behavior as a function of , reflecting the impact of ZnFe2O3 doping on the dielectric properties of the composite. For the dielectric constant (ε1), ZIF-8@ZnFeO3 exhibits higher values across the measured wavelength range than pristine ZIF-8. This enhancement can be attributed to the introduction of ZnFe2O3 nanoparticles, which increase the polarization within the composite. The peak observed in the higher wavelength region for ZIF-8@ZnFe2O3 signifies improved charge storage capability due to the synergistic effect between the ZnFe2O3 and the ZIF-8 framework. The dielectric loss (ε2) spectra reveal the energy dissipation associated with the charge migration and relaxation processes. ZIF-8@ZnFe2O3 demonstrates a sharper peak at shorter wavelengths, indicating a higher degree of energy absorption, which is characteristic of materials with enhanced dipole relaxation mechanisms. The smoother behavior of ε2 at longer wavelengths reflects reduced loss, aligning with improved dielectric performance.
The enhanced dielectric behavior of ZIF-8@ZnFe2O3 is consistent with findings reported in the literature for metal oxide-doped MOF composites. Studies have shown that the incorporation of metal oxides such as ZnFe2O3 increases the dielectric constant and loss due to improved electronic coupling and the formation of heterogeneous interfaces, which facilitate charge polarization [37]. Additionally, the synergistic effects between metal oxides and MOFs have been highlighted as a key factor in achieving enhanced dielectric performance [38]. Overall, the improved dielectric constant and reduced energy loss of ZIF-8@ZnFe2O3 underscore its potential for energy storage and electronic applications. These enhancements are indicative of the structural and compositional advantages introduced by ZnFe2O3 doping, making this composite material a promising candidate for next-generation dielectric and optoelectronic devices.

2.3. Photocatalytic Study

To better understand how different amounts of dopant affect photocatalytic efficiency, we compared 5 wt% and 10 wt% ZnFe2O3-doped ZIF-8 composites directly. As illustrated in the newly added Figure 10, all samples exhibited a significant decrease in MB absorbance after 60 min under visible light, confirming photocatalytic activity. The ZnFe2O3@ZIF-8 (10 wt%) composite exhibited the most significant reduction in MB absorbance, outperforming both the 5 wt% sample and pristine ZIF-8. This improvement is attributed to having more catalytic sites and better charge transfer with increased doping. The data clearly show that the 10 wt% ZnFe2O3-doped ZIF-8 composite works better as a photocatalyst than the 5 wt% version.
The photocatalytic efficacy of ZIF-8 and ZIF-8@ZnFe2O3 was assessed by observing the degradation of a methylene blue (MB) solution under visible light at different time intervals, as shown in Figure 11. Figure 11a shows the absorbance of MB in the presence of ZIF-8. The initial absorbance of MB is significant, but it decreases steadily as visible light increases. At 0 min, the absorbance is approximately 0.9, which reduces to around 0.6 after 30 min of exposure to visible light. The absorbance decreases, reaching around 0.4 after 60 min and 0.2 after 120 min. This indicates that the ZIF-8 catalyst effectively degrades MB over time. In Figure 11b, the absorbance of MB is shown in the presence of ZIF-8@ZnFe2O3. The trend is like that observed in Figure 11a, but the degradation of MB is more pronounced in this case. The initial absorbance is similar but rapidly drops with increasing visible-light time. After 30 min, the absorbance decreases to approximately 0.4; after 60 min, it reaches around 0.2.
The absorbance continues to decline, reaching around 0.1 after 120 min. This suggests that the ZIF-8@ZnFe2O3 catalyst is more effective at degrading MB than the ZIF-8 catalyst. Overall, these results demonstrate the potential of both ZIF-8 and ZIF-8@ZnFe2O3 catalysts for the degradation of methylene blue under visible-light conditions, with the latter showing improved performance. The superior photocatalytic performance of ZIF-8@ZnFe2O3 can be attributed to the synergistic effects between ZnFe2O3 and the ZIF-8 framework. The incorporation of ZnFe2O3 enhances the light absorption capacity of the composite, as demonstrated by its higher refractive index and broader optical response. Additionally, ZnFe2O3 facilitates improved charge separation and transfer at the interface, thereby reducing electron–hole recombination and increasing the availability of reactive oxygen species for the degradation of MB. The faster degradation of MB in the presence of ZIF-8@ZnFe2O3 highlights its potential as a highly efficient photocatalyst, making this composite material a promising candidate for wastewater treatment and environmental remediation applications.
Figure 12 illustrates the photocatalytic degradation of MB under visible light in the presence of different catalysts: pure MB solution (without catalyst), ZIF-8, and ZIF-8@ZnFe2O3 composites. The MB concentration exhibited only a minimal decrease in the absence of any catalyst, indicating negligible photolysis under visible light and underscoring the necessity of a photocatalyst for effective degradation. In contrast, the inclusion of ZIF-8 led to a significant reduction in MB concentration over time, demonstrating its inherent photocatalytic activity. Notably, ZIF-8@ZnFe2O3 exhibited superior performance, with the steepest decline in MB concentration, particularly during the initial stages of visible light.
The enhanced performance of the ZIF-8@ZnFe2O3 composite can be attributed to the synergistic interaction between ZIF-8 and ZnFe2O3. ZIF-8 provides a high surface area and serves as a light-harvesting framework, while ZnFe2O3 contributes to improved charge carrier dynamics. Specifically, the incorporation of ZnFe2O3 enhances charge separation and light absorption, likely contributing to the improved photocatalytic efficiency observed in the composite material. Similar findings have been reported in studies where doping metal–organic frameworks (MOFs) with metal oxides, such as ZnO, enhances photocatalytic activity due to improved charge separation and transfer [39]. Furthermore, combining ZIF-8 with metal oxides results in molecule size selectivity and excellent photocatalytic efficiency [40]. This observation aligns with studies demonstrating that coupling metal oxides with MOFs enhances photocatalytic efficiency under both UV and visible light. For instance, a novel nanocomposite of ZIF-8 and MoSe2 has been shown to exhibit enhanced photocatalytic activity in degrading organic pollutants [41]. The lack of substantial degradation in the absence of a catalyst underscores the necessity of a photocatalyst for efficient MB degradation. While ZIF-8 alone shows moderate activity due to its structural and optical properties, the integration of ZnFe2O3 significantly amplifies its photocatalytic potential, making the ZIF-8@ZnFe2O3 composite a promising candidate for environmental remediation applications.
Photocatalytic efficiency (PCE%) was calculated using the formula
P C E % = C o C t C o × 100 = 1 A t A o × 100
where Ao is the initial absorbance and At is the absorbance at time t. The PCE% of ZIF-8 and ZIF-8@ZnFe2O3 catalysts in degrading methylene blue (MB) under visible light is presented in Figure 13. The plot shows the PCE% of each catalyst against UV light irradiation time (in minutes). The results demonstrate that ZIF-8 and ZIF-8@ZnFe2O3 catalysts exhibit significant photocatalytic activity towards MB degradation under UV light. However, the ZIF-8@ ZnFe2O3 catalyst displays a noticeably higher PCE than ZIF-8 throughout the visible-light period. The PCE% of ZIF-8 increases steadily with increasing visible-light time, reaching a maximum efficiency of around 60% after 120 min.
In contrast, the PCE% of ZIF-8@ZnFe2O3 increases more rapidly, achieving maximum efficiency of approximately 90% after 120 min of visible light exposure. The enhanced photocatalytic performance of ZIF-8@ZnFe2O3 can be attributed to the synergistic effect of the ZnFe2O3 component, which improves the separation and transfer of photogenerated electrons and holes, thereby increasing the catalytic activity. The higher PCE% of ZIF-8@ZnFe2O3 also suggests that the ZnFe2O3 component may have increased the surface area, pore volume, and/or adsorption capacity of the catalyst, allowing for more efficient degradation of MB. These findings are consistent with previous studies. For instance, Jing et al. [42] demonstrated that ZIF-8 exhibits efficient photocatalytic activity for MB degradation under visible light, which is attributed to its ability to generate hydroxyl radicals. The enhanced photocatalytic efficiency observed in the ZIF-8@ZnFe2O3 composite highlights the potential of integrating metal–organic frameworks (MOFs) with metal oxides to develop effective photocatalysts for environmental remediation applications.
The disintegration rate of a pseudo-first-order response obeyed the kinetic expression [43]:
l n C t C o = k K t =   k a p p   t
where k is the degradation rate constant, K is the adsorption equilibrium constant, and kapp is the apparent kinetic rate constant. The Ct/Co can be replaced by At/Ao because the concentricity is specific to the absorbance. Figure 14 illustrates the plot of a linear correlation between ln(Ct/Co) and visible-light time, indicating that the degradation of MB follows a pseudo-first-order kinetic model. The linear fitting of the data suggests that the degradation rate of MB is directly proportional to the concentration of MB. The slope of the linear fit represents the apparent rate constant (kapp) of the degradation reaction.
Comparing the two catalysts, the slope of the linear fit for ZIF-8@ZnFe2O3 is steeper than that of ZIF-8, indicating a higher apparent rate constant and a faster degradation rate of MB in the presence of ZIF-8@ZnFe2O3. This finding is consistent with previous results, which demonstrated that ZIF-8@ZnFe2O3 exhibits higher photocatalytic efficiency and degradation rates compared to ZIF-8. The higher degradation rate of MB in the presence of ZIF-8@ZnFe2O3 can be attributed to the enhanced photocatalytic activity of the ZIF-8@ZnFe2O3 catalyst, which is due to the synergistic effect of the ZnFe2O3 component. Incorporating ZnFe2O3 into the ZIF-8 framework may have increased the surface area, pore volume, and/or adsorption capacity of the catalyst, leading to more efficient degradation of MB.
Finally, the photocatalytic performance of various catalysts for the degradation of methylene blue (MB) under different conditions has been extensively reported in the literature. Table 1 summarizes the photocatalytic efficiency (PCE%) and apparent rate constants (kₐₚₚ) for ZnFe2O3-doped ZIF-8 and other catalysts, highlighting their strengths and limitations. Pristine ZIF-8 exhibits moderate photocatalytic activity among the studied catalysts, attributed to its structural and optical properties. However, its performance is limited by rapid electron–hole recombination and a wide bandgap. Integrating ZnFe2O3 into the ZIF-8 framework significantly improves its photocatalytic efficiency, achieving a degradation efficiency of 90% under UV light with an apparent rate constant of 12.51 × 10−3 min−1. This enhancement can be attributed to the synergistic effects between ZnFe2O3 and ZIF-8, which improve charge separation, reduce recombination rates, and enhance light absorption.
In comparison, other catalysts [39,40,41,42], such as Fe/ZnO/SiO2 nanoparticles, have demonstrated 100% MB degradation efficiency under visible light. These nanoparticles benefit from their ability to utilize a broader light spectrum, but their synthesis and stability can present challenges for large-scale applications. Similarly, MXene-based photocatalysts have shown high efficiencies, reaching 90.2%, due to their excellent electrical conductivity and surface activity. However, their stability and long-term performance under various environmental conditions remain to be investigated. The results from this study position ZnFe2O3-doped ZIF-8 as a competitive and versatile photocatalyst for environmental remediation, particularly under visible-light conditions. The improved performance underscores the benefits of integrating metal–organic frameworks (MOFs) with metal oxides, offering a pathway for developing advanced materials for wastewater treatment and other photocatalytic applications.

2.4. Photocatalytic Reusability and Stability

The long-term stability and reusability of photocatalysts are crucial parameters for evaluating their feasibility in practical environmental remediation applications. To assess the photocatalytic durability of the prepared materials, cyclic degradation experiments were conducted for five consecutive runs using methylene blue (MB) under visible light, as shown in Figure 15.
The ZnFe2O3@ZIF-8 composite maintained a high photocatalytic efficiency throughout the five cycles, with only a minor reduction from ~90% in the first cycle to ~86% in the fifth. This minimal decrease (~4%) indicates strong photostability, robust structural integrity, and negligible catalyst leaching or deactivation. In contrast, pristine ZIF-8 exhibited a more noticeable decline, with the photocatalytic efficiency dropping from ~68% to ~63% by the fifth cycle. The superior recyclability of ZnFe2O3@ZIF-8 is attributed to several factors: (i) the strong interaction between ZnFe2O3 nanoparticles and the ZIF-8 framework, which inhibits photocorrosion and enhances charge separation; (ii) the structural stability of the MOF during light-induced catalytic reactions; and (iii) the retention of high surface area and active sites after repeated use. These findings align with prior studies, which demonstrate that MOF/metal oxide hybrids often exhibit better cyclic performance compared to pristine MOFs due to synergistic interfacial effects [41,42]. Overall, the excellent recyclability of ZnFe2O3@ZIF-8 under visible-light illumination emphasizes its potential as a robust and reusable photocatalyst for practical wastewater treatment systems.

2.5. Mechanism of Photocatalytic Degradation

The enhanced photocatalytic activity of ZnFe2O3@ZIF-8 under visible light is primarily attributed to the synergistic interaction between ZnFe2O3 nanoparticles and the ZIF-8 framework, which promotes efficient charge separation, reduced electron-hole recombination, and generation of reactive species (Figure 16). Upon visible-light irradiation (λ ≥ 420 nm), both ZIF-8 and ZnFe2O3 absorb photons and generate electron–hole pairs (e/h+):
Z n F e 2 O 3 + h ν e + h +
Due to the heterojunction formed between ZIF-8 and ZnFe2O3, photogenerated electrons from ZnFe2O3 can transfer to the conduction band of ZIF-8, while holes remain in the valence band of ZnFe2O3. This spatial separation suppresses recombination and allows the chargers to participate in redox reactions at the catalyst surface:
Oxygen reduction: Electrons reduce dissolved oxygen into superoxide radicals:
O 2 + e O 2
Hydroxyl radical formation: Superoxide radicals and holes further generate hydroxyl radicals:
O 2 + H + H O 2 O H
H 2 O + h + O H + H +
Organic dye degradation: These reactive oxygen species (⋅OH,⋅O2) attack the methylene blue dye molecules and oxidize them into harmless products:
M B + O 2 C O 2 + H 2 O + o t h e r   p r o d u c t s
Thus, the heterojunction structure facilitates enhanced separation and migration of charge carriers, improving the production of active species for dye degradation.

3. Experimental Technique

3.1. Sample Fabrication

The synthesis of ZnFe2O3-doped ZIF-8 is achieved through a solvothermal method. First, we dissolve 2.97 g (10 mmol) of zinc nitrate hexahydrate (Zn(NO3)2·6H2O) and 3.28 g (40 mmol) of 2-methylimidazole (2-MIM) in 50 mL of deionized water, stirring at 500 rpm for 30 min. The obtained solution showed ZIF-8. This method of making ZIF-8 aligns with earlier techniques that also employ room-temperature precipitation, as described by Park et al. [48]. Separately, ZnFe2O3 nanoparticles were prepared via a thermal decomposition method [22]. The Zn/MIM molar ratio was selected as 1:4 to ensure optimal nucleation and crystal growth, consistent with similar excess-ligand approaches reported in the literature for improved ZIF-8 morphology. The ZnFe2O3 nanoparticles (10 wt%) were synthesized by thermally decomposing a mixed nitrate solution containing Zn(NO3)2·6H2O and Fe(NO3)3·9H2O in a 1:2 molar ratio at 450 °C, following the protocol from our earlier work [22]. The mixture was transferred to a Teflon-lined autoclave and heated at 120 °C for 24 h, allowing the nucleation and growth of the ZIF-8 framework while incorporating ZnFe2O3. The autoclave was allowed to cool to room temperature. The resultant product was collected by centrifugation at 5000 rpm, washed three times with ethanol and deionized water, and then dried in a vacuum oven at 60 °C for 12 h to obtain the ZnFe2O3-doped ZIF-8 powders. For comparison, pristine ZIF-8 was synthesized under identical conditions without the addition of ZnFe2O3.
Thin films of ZIF-8 and ZnFe2O3-doped ZIF-8 can be prepared via a spin-coating method, which provides a uniform layer suitable for optical characterization. Glass slides were used as substrates. The cleaning process involved ultrasonic cleaning in acetone and ethanol (each for 10 min), followed by rinsing with deionized water and blow-drying with nitrogen to ensure surface cleanliness and uniformity. The synthesized ZIF-8 or ZnFe2O3-doped ZIF-8 powder (20 mg) was dispersed in a mixture of polyvinyl alcohol (1 wt.%) and ethanol (10 mL) to form a slurry. The mixture was sonicated to obtain homogeneous dispersion. The resulting slurry was deposited onto the prepared substrates using a spin coater. The spin-coating process was performed at 2000 rpm for 30 s, ensuring a uniform film thickness. After deposition, the coated film was allowed to dry at room temperature for 12 h to remove residual solvent, followed by thermal curing at 60 °C for 2 h to enhance adherence and stability.

3.2. Measurement Tools

X-ray diffraction (XRD) patterns were recorded using a PANalytical Empyrean X-ray Diffractometer (Almelo, The Netherlands) with CuKα radiation (λ = 1.5406 Å). The data were collected over a 2θ range of 5–80° at a scan rate of 0.02 °/s. Thermogravimetric analysis (TGA) and differential thermal analysis (DTA) were performed using a TGA 5500 Thermogravimetric Analyzer (TA Instruments, New Castle, DE, USA) under a nitrogen atmosphere at a heating rate of 10 °C/min, from room temperature to 800 °C. Scanning electron microscopy (SEM) images were obtained using a JEOL JSM-7600F field-emission SEM (Tokyo, Japan) at an accelerating voltage of 5 kV. Fourier transform infrared (FTIR) spectroscopy was performed using a Bruker Tensor 27 spectrometer (Bremen, Germany) over a wavenumber range of 4000–400 cm−1 to analyze functional groups and bonding interactions. A Shimadzu UV-3600 UV–visible spectrophotometer (Kyoto, Japan) was used.

3.3. Photocatalytic Degradation

The photocatalytic activity of ZIF-8 and ZIF-8 (10 wt%) incorporated with ZnFe2O3 (ZIF-8@ZnFe2O3) was evaluated using methylene blue (MB) dye degradation under visible light. A 25 mL solution of MB (10 mg/L) was prepared in deionized water and used as the model pollutant. For the experiments, 30 mg of the catalyst was dispersed into the MB solution and stirred in the dark for 30 min to achieve adsorption–desorption equilibrium. The solution was then exposed to visible light from a 300 W xenon lamp (λ ≥ 420 nm, and an AM 1.5G optical filter. Aliquots were withdrawn at 10 min intervals, and the remaining MB concentration was determined using a UV–vis spectrophotometer (Jenway, 6800, Staffordshire, UK) at 664 nm.

4. Conclusions

This study successfully synthesized 10 wt% ZnFe2O3-doped ZIF-8 composites using a solvothermal method, significantly enhancing their structural, optical, and photocatalytic properties. The incorporation of ZnFe2O3 nanoparticles into the ZIF-8 framework was confirmed through XRD, FTIR, and SEM analyses, which revealed the preservation of the crystalline structure and the formation of a porous composite with well-distributed ZnFe2O3 nanoparticles. However, it is essential to note that the structural identification of ZIF-8 in the doped composites is not yet complete. The XRD patterns, especially at high dopant loading (10 wt%), are very different from those of regular ZIF-8. For example, there are sharp diffraction peaks that come from crystalline ZnFe2O3. These data imply that the structure is partially distorted and the phases are separated, which means that ZIF-8’s phase purity and crystallographic integrity are not satisfactory at high dopant concentrations. The optical studies demonstrated a reduction in the bandgap from 4.37 eV (pristine ZIF-8) to 4.26 eV for the doped composite, along with improved light absorption and extinction coefficients, highlighting the material’s enhanced light-harvesting capabilities. The photocatalytic performance of the composite was significantly superior, achieving a 90% degradation efficiency of methylene blue under visible light compared to 67.8% for pristine ZIF-8. This enhancement is attributed to the synergistic effects between ZnFe2O3 and ZIF-8, which improve charge carrier separation, reduce electron–hole recombination, and facilitate the generation of reactive oxygen species. These findings demonstrate the potential of ZnFe2O3-doped ZIF-8 composites as highly efficient photocatalysts for environmental remediation, particularly in wastewater treatment. This study provides a foundation for further exploration of ZnFe2O3-doped ZIF-8 composites in advanced energy and environmental applications. Future work could optimize the synthesis process, explore visible-light-driven photocatalytic performance, and scale up the material for industrial applications. Additionally, integrating this composite into multifunctional energy storage systems and its potential for broader environmental applications merits further investigation.

Author Contributions

F.A., funding acquisition, formal analysis, writing—original draft. T.H., conceptualization, methodology, formal analysis, writing—original draft. H.A.-G., data curation, writing—review and editing. B.A., methodology, formal analysis. A.D., methodology, data curation, writing—review and editing. All authors have read and agreed to the published version of the manuscript.

Funding

Deanship of Scientific Research and Libraries at Princess Nourah bint Abdulrahman University the Research Group project, Grant No. (RG-1445-0030).

Data Availability Statement

The original contributions presented in the study are included in the article; further inquiries can be directed to the corresponding author.

Acknowledgments

The authors extend their appreciation to the Deanship of Scientific Research and Libraries at Princess Nourah bint Abdulrahman University for funding this research work through the Research Group project, Grant No. (RG-1445-0030).

Conflicts of Interest

The authors declare that they have no competing interests or personal relationships that could be perceived as influencing the work reported in this study.

References

  1. Mane, P.V.; Rego, R.M.; Yap, P.L.; Losic, D.; Kurkuri, M.D. Unveiling cutting-edge advances in high-surface-area porous materials for the efficient removal of toxic metal ions from water. Prog. Mater. Sci. 2024, 146, 101314. [Google Scholar] [CrossRef]
  2. Jia, T.; Gu, Y.; Li, F. Progress and potential of metal-organic frameworks (MOFs) for gas storage and separation: A review. J. Environ. Chem. Eng. 2022, 10, 108300. [Google Scholar] [CrossRef]
  3. Hu, S.; Guo, G.; Zhang, J.; Khan, M.N.; Xu, S.; Yang, F.; Schwandt, B.W.; Hu, Z.; Zou, J. Advanced porous MOF materials and technologies for high-efficiency ppm-level toxic gas separation. Mater. Sci. Eng. R 2024, 161, 100874. [Google Scholar] [CrossRef]
  4. Zhang, Y.; Yu, X.; Hou, Y.; Liu, C.; Xie, G.; Chen, X. Current research status of MOF materials for catalysis applications. Mol. Catal. 2024, 555, 113851. [Google Scholar] [CrossRef]
  5. Qu, L.; Xu, Y.; Cui, W.; Wu, L.; Feng, Y.; Gu, Y.; Pan, H. Trends in conductive MOFs for sensing: A review. Anal. Chim. Acta 2024, 1336, 343307. [Google Scholar] [CrossRef] [PubMed]
  6. Gargiulo, V.; Alfè, M.; Giordano, L.; Lettieri, S. Materials for Chemical Sensing: A Comprehensive Review on the Recent Advances and Outlook Using Ionic Liquids, Metal-Organic Frameworks (MOFs), and MOF-Based Composites. Chemosensors 2022, 10, 290. [Google Scholar] [CrossRef]
  7. Zhang, X.; Zhang, S.; Tang, Y.; Huang, X.; Pang, H. Recent advances and challenges of metal-organic framework/graphene-based composites. Compos. Part B Eng. 2022, 230, 109532. [Google Scholar] [CrossRef]
  8. Qi, X.; Wang, H.; Huang, B.; Yuan, X.; Zhang, J.; Zhang, J.; Jiang, L.; Xiong, T.; Zeng, G. State-of-the-art advances and challenges of iron-based metal-organic frameworks from attractive features, synthesis to multifunctional applications. Small 2019, 15, 1803088. [Google Scholar]
  9. Ghafari, S.; Kazemzad, M.; Naderi, N.; Eshraghi, M.J. Performance Improvement of Photodetectors Based on ZIF-8 Nanostructures on Porous Silicon Substrate. J. Electron. Mater. 2024, 53, 1577–1589. [Google Scholar] [CrossRef]
  10. Sabzehmeidani, M.M.; Gafari, S.; Jamali, S.; Kazemzad, M. Concepts; fabrication, and applications of MOF thin films in optoelectronics: A review. Appl. Mater. Today 2024, 38, 102153. [Google Scholar] [CrossRef]
  11. Teng, D.; Zheng, W.; Wu, L.; Murtaza, G.; Meng, Z.; Qiu, L. Smart manufacturing of ZIF-8 photonic crystals: Catalytic formaldehyde degradation and colorful eco-friendly coatings. Chem. Eng. J. 2024, 495, 15352. [Google Scholar] [CrossRef]
  12. Karim, M.R.; Choi, C.-H.; Mohammad, A.; Yoon, T. Fabrication of high-performance supercapacitor of surface-engineered ZIF-8 for energy storage applications. J. Energy Storage 2024, 93, 112199. [Google Scholar] [CrossRef]
  13. Chafiq, M.; Chaouiki, A.; Al-Moubaraki, A.H.; Ko, Y.G. Recent progress in ZIF nanocomposite materials for wastewater pollutant in aqueous solution: A mini-review. Process Saf. Environ. Prot. 2024, 184, 1017–1033. [Google Scholar] [CrossRef]
  14. Li, K.; Miwornunyuie, N.; Chen, L.; Jingyu, H.; Amaniampong, P.S.; Koomson, D.A.; Ewusi-Mensah, D.; Xue, W.; Li, G.; Lu, H. Sustainable Application of ZIF-8 for Heavy-Metal Removal in Aqueous Solutions. Sustainability 2021, 13, 984. [Google Scholar] [CrossRef]
  15. Dai, H.; Yuan, X.; Jiang, L.; Wang, H.; Zhang, J.; Zhang, J.; Xiong, T. Recent advances on ZIF-8 composites for adsorption and photocatalytic wastewater pollutant removal: Fabrication, applications and perspective. Coord. Chem. Rev. 2021, 441, 213985. [Google Scholar] [CrossRef]
  16. Shen, B.; Wang, B.; Zhu, L.; Jiang, L. Properties of Cobalt- and Nickel-Doped Zif-8 Framework Materials and Their Application in Heavy-Metal Removal from Wastewater. Nanomaterials 2020, 10, 1636. [Google Scholar] [CrossRef]
  17. Kaneti, Y.V.; Moriceau, J.; Liu, M.; Yuan, Y.; Zakaria, Q.; Jiang, X.; Yu, A. Hydrothermal synthesis of ternary α-Fe2O3–ZnO–Au nanocomposites with high gas-sensing performance. Sens. Actuators B Chem. 2015, 209, 889–897. [Google Scholar] [CrossRef]
  18. Lee, Y.-S.; Kim, H.-T.; Yoo, K.-O. Effect of ferric oxide on the high-temperature removal of hydrogen sulfide over ZnO-Fe2O3 mixed metal oxide Sorbent. Ind. Eng. Chem. Res. 1995, 34, 1181–1188. [Google Scholar] [CrossRef]
  19. Alsharif, M.A.; Darwish, A.A.A.; Alghamdi, N.; Alfadhli, S.; Khasim, S.; Ahmed, S.; Hamdalla, T.A. Synergistic Enhancements of Zn-ZIF with Nano Zinc Oxide for Hydrogen adsorption, energy storage, and photocatalytic technologies. Ceram. Int. 2024, 50, 47677–47686. [Google Scholar] [CrossRef]
  20. Saravanan, M.; Sabari Girisun, T.C. Enhanced nonlinear optical absorption and optical limiting properties of superparamagnetic spinel zinc ferrite-decorated reduced graphene oxide nanostructures. Appl. Surf. Sci. 2017, 392, 904–991. [Google Scholar]
  21. Danilo, D.; Calvete, M.J.F.; Hanack, M. Nonlinear optical materials for the smart filtering of optical radiation. Chem. Rev. 2016, 116, 13043–13233. [Google Scholar] [CrossRef]
  22. Saba, S.; Alsharari, A.M.; Aldaleeli, N.Y.; Aljohani, M.M.; Hamdalla, T.A. Structural, Electrochemical, and Supercapacitor Characterization of Double Metal Oxides Doped Within ZIF-8 Composites. Processes 2025, 13, 859. [Google Scholar] [CrossRef]
  23. Liao, W.-L.; Abdelaal, M.M.; Amirtha, R.-M.; Fang, C.-C.; Yang, C.-C.; Hung, T.-F. In Situ Construction of Nitrogen-Doped and Zinc-Confined Microporous Carbon Enabling Efficient Na+-Storage Abilities. Int. J. Mol. Sci. 2023, 24, 8777. [Google Scholar] [CrossRef]
  24. Cravillon, J.; Schröder, C.A.; Bux, H.; Rothkirch, A.; Caro, J.; Wiebcke, M. Formate modulated solvothermal synthesis of ZIF-8 investigated using time-resolved in situ X-ray diffraction and scanning electron microscopy. CrystEngComm 2012, 14, 492–498. [Google Scholar] [CrossRef]
  25. Zappa, D.; Galstyan, V.; Kaur, N.; Arachchige, H.M.M.M.; Sisman, O.; Comini, E. “Metal oxide-based heterostructures for gas sensors”—A review. Anal. Chim. Acta 2018, 1039, 1–23. [Google Scholar] [CrossRef] [PubMed]
  26. Yu, J.; Yang, H.; Li, X.; Chen, X.; Li, W.; Li, X.; Zhang, C. MOF-on-MOF-derived hollow tubular ZnFe2O4/Er-ZnO heterojunctions for the efficient degradation of tetracycline by visible light photocatalysis. Appl. Surf. Sci. 2025, 684, 161634. [Google Scholar] [CrossRef]
  27. Dutta, T.; Kim, T.; Vellingiri, K.; Tsang, D.C.W.; Shon, J.R.; Kim, K.-H.; Kumar, S. Recycling and regeneration of carbonaceous and porous materials through thermal or solvent treatment. Chem. Eng. J. 2019, 364, 514–529. [Google Scholar] [CrossRef]
  28. Lei, Y.; Xiao, Y.; Chen, X.; Zhang, W.; Yang, X.; Yang, H.; Fang, D. Research Progress on CO2 Capture and Catalytic Conversion of Metal-Organic Frameworks Materials. Catalysts 2025, 15, 421. [Google Scholar] [CrossRef]
  29. Chen, P.; Wang, S.; Peng, H.; Wang, S.; Wang, J.; Zhao, X.; Chen, C.; Zhou, H. Preparation of ZIF-8/Fe3O4 modified halloysite nanotubes blended ultrafiltration membranes via magnetic field regulation: Enhanced performance and efficient filler distribution. Sep. Purif. Technol. 2025, 362, 131833. [Google Scholar] [CrossRef]
  30. Jiao, L.; Seow, J.Y.; Skinner, W.S.; Wang, Z.U.; Jiang, H.L. Metal-organic frameworks: Structures and functional applications. Mater. Today 2019, 27, 43–68. [Google Scholar] [CrossRef]
  31. El-Nahass, M.M.; Zeyada, H.M.; Abd-El-Rahman, K.F.; Farag, A.A.M.; Darwish, A.A.A. Fourier-transform infrared and optical absorption spectra of 4-tricyanovinyl-N,N-diethylaniline thin films. Spectrochim. Acta Part A Mol. Biomol. Spectrosc. 2008, 69, 205–210. [Google Scholar] [CrossRef]
  32. Guan, D.; Xu, H.; Huang, Y.-C.; Jing, C.; Tsujimoto, Y.; Xu, X.; Lin, Z.; Tang, J.; Wang, Z.; Sun, X.; et al. Operando Studies Redirect Spatiotemporal Restructuration of Model Coordinated Oxides in Electrochemical Oxidation. Adv. Mater. 2025, 37, 2413073. [Google Scholar] [CrossRef]
  33. El-Zaidia, E.F.M.; Ali, H.A.M.; Hamdalla, T.A.; Darwish, A.A.A.; Hanafy, T.A. Optical linearity and bandgap analysis of Erythrosine B doped in polyvinyl alcohol films. Opt. Mater. 2020, 100, 109661. [Google Scholar] [CrossRef]
  34. Nafee, S.S.; Hamdalla, T.A.; Darwish, A.A.A. Studies of the morphology and optical properties of nano erbium oxide embedded in PMMA matrix. Opt. Laser Technol. 2020, 129, 106282. [Google Scholar] [CrossRef]
  35. Yang, X.; Wen, Z.; Wud, Z.; Luo, X. Synthesis of ZnO/ZIF-8 hybrid photocatalysts derived from ZIF-8 with enhanced photocatalytic activity. Inorg. Chem. Front. 2018, 5, 687. [Google Scholar] [CrossRef]
  36. Sen, P.; Bhattacharya, P.; Mukherjee, G.; Ganguly, J.; Marik, B.; Thapliyal, D.; Verma, S.; Verros, G.D.; Chauhan, M.S.; Arya, R.K. Advancements in Doping Strategies for Enhanced Photocatalysts and Adsorbents in Environmental Remediation. Technologies 2023, 11, 144. [Google Scholar] [CrossRef]
  37. Zhang, Y.; Lin, Y.; Duan, T.; Song, L. Interfacial engineering of heterogeneous catalysts for electrocatalysis. Mater. Today 2021, 48, 115–134. [Google Scholar] [CrossRef]
  38. Lv, Y.; Tian, J.; Chen, Z.; Wang, J.; Ma, L.-A.; Zhang, L.; Chen, S.; Wang, Q.; Ye, X. MXene/bimetallic CoNi-MOF-derived magnetic-dielectric balanced composites with multiple heterogeneous interfaces for excellent microwave absorption. Chem. Eng. J. 2023, 478, 147413. [Google Scholar] [CrossRef]
  39. Wang, X.; Liu, J.; Leong, S.; Lin, X.; Kong, J.W.B.; Xu, Y.; Low, Z.-X.; Yao, J.; Wang, H. Rapid Construction of ZnO@ZIF-8 Heterostructures with Size-Selective Photocatalysis Properties. ACS Appl. Mater. Interfaces 2016, 8, 9080–9087. [Google Scholar] [CrossRef] [PubMed]
  40. Elaouni, A.; El Ouardi, M.; Zbair, M.; BaQais, A.; Saadi, M.; Ahsaine, H.A. ZIF-8 metal-organic framework materials as a superb platform for the removal and photocatalytic degradation of organic pollutants: A review. RSC Adv. 2022, 12, 31801–31817. [Google Scholar] [CrossRef]
  41. Mittal, H.; Ivaturi, A. MoSe2-modified ZIF-8 novel nanocomposite for photocatalytic remediation of textile dye and antibiotic-contaminated wastewater. Environ. Sci. Pollut. Res. 2023, 30, 4151–4165. [Google Scholar] [CrossRef] [PubMed]
  42. Jing, H.-P.; Wang, C.-C.; Zhang, Y.-W.; Wang, P.; Li, R. Photocatalytic degradation of methylene blue in ZIF-8. RSC Adv. 2014, 4, 54454–54462. [Google Scholar] [CrossRef]
  43. Rashad, M.; Hamdalla, T.A.; Darwish, A.A.A.; Seleim, S.M. High-performance efficiency of water purification using Cr2O3 nanoparticles synthesized by thermal combustion technique. Mater. Res. Express 2019, 6, 065048. [Google Scholar] [CrossRef]
  44. Mohamed, R.M.; Mkhalid, I.A.; Baeissa, E.S.; Al-Rayyani, M.A. Photocatalytic Degradation of Methylene Blue by Fe/ZnO/SiO2 Nanoparticles under visible light. J. Nanotechnol. 2012, 2012, 329082. [Google Scholar] [CrossRef]
  45. Orasugh, J.T.; Temane, L.T.; Ray, S.S. Application of MXenes in Water Purification, CO2 Capture and Conversion. In Two-Dimensional Materials for Environmental Applications; Springer International Publishing: Cham, Switzerland, 2023; pp. 17–74. [Google Scholar]
  46. Rahman, I.A.; Ayob, M.T.M.; Radiman, S. Enhanced Photocatalytic Performance of NiO-Decorated ZnO Nanowhiskers for Methylene Blue Degradation. J. Nanotechnol. 2014, 2014, 212694. [Google Scholar] [CrossRef]
  47. Chen, M.L.; Cho, K.Y.; Oh, W.C. Synthesis and photocatalytic behaviors of Cr2O3–CNT/TiO2 composite materials under visible light. J. Mater. Sci. 2010, 45, 6611–6616. [Google Scholar] [CrossRef]
  48. Park, K.S.; Ni, Z.; Côté, A.P.; Choi, J.Y.; Huang, R.; Uribe-Romo, F.J.; Chae, H.K.; O’Keeffe, M.; Yaghi, O.M. Exceptional chemical and thermal stability of zeolitic imidazolate frameworks. Proc. Natl. Acad. Sci. USA 2006, 103, 10186–10191. [Google Scholar] [CrossRef]
Figure 1. XRD patterns of pristine ZIF-8, 5 wt% ZnFe2O3@ZIF-8 [22], and 10 wt% ZnFe2O3@ZIF-8 composites.
Figure 1. XRD patterns of pristine ZIF-8, 5 wt% ZnFe2O3@ZIF-8 [22], and 10 wt% ZnFe2O3@ZIF-8 composites.
Catalysts 15 00720 g001
Figure 2. Thermal analysis (TGA and DTA) for ZIF-8 (10 wt%) incorporated with ZnFe2O3.
Figure 2. Thermal analysis (TGA and DTA) for ZIF-8 (10 wt%) incorporated with ZnFe2O3.
Catalysts 15 00720 g002
Figure 3. SEM images for (a) ZIF-8 and (b) ZIF-8 (10 wt%) incorporated with ZnFe2O3.
Figure 3. SEM images for (a) ZIF-8 and (b) ZIF-8 (10 wt%) incorporated with ZnFe2O3.
Catalysts 15 00720 g003
Figure 4. FTIR analysis for ZIF-8 and ZIF-8 (10 wt%) incorporated with ZnFe2O3.
Figure 4. FTIR analysis for ZIF-8 and ZIF-8 (10 wt%) incorporated with ZnFe2O3.
Catalysts 15 00720 g004
Figure 5. Optical transmission and reflection of ZIF-8 and ZIF-8 (10 wt%) incorporated with ZnFe2O3.
Figure 5. Optical transmission and reflection of ZIF-8 and ZIF-8 (10 wt%) incorporated with ZnFe2O3.
Catalysts 15 00720 g005
Figure 6. Extinction coefficient (k) spectra for ZIF-8 and ZIF-8 (10 wt%) incorporated with ZnFe2O3.
Figure 6. Extinction coefficient (k) spectra for ZIF-8 and ZIF-8 (10 wt%) incorporated with ZnFe2O3.
Catalysts 15 00720 g006
Figure 7. Tauc plots for indirect energy gap calculation of ZIF-8 and ZIF-8 (10 wt%) incorporated with ZnFe2O3.
Figure 7. Tauc plots for indirect energy gap calculation of ZIF-8 and ZIF-8 (10 wt%) incorporated with ZnFe2O3.
Catalysts 15 00720 g007
Figure 8. Refractive index (n) spectra for ZIF-8 and ZIF-8 (10 wt%) incorporated with ZnFe2O3.
Figure 8. Refractive index (n) spectra for ZIF-8 and ZIF-8 (10 wt%) incorporated with ZnFe2O3.
Catalysts 15 00720 g008
Figure 9. (a) Dielectric constant (ε1) and (b) dielectric loss (ε2) spectra of ZIF-8 and ZIF-8 (10 wt%) incorporated with ZnFe2O3.
Figure 9. (a) Dielectric constant (ε1) and (b) dielectric loss (ε2) spectra of ZIF-8 and ZIF-8 (10 wt%) incorporated with ZnFe2O3.
Catalysts 15 00720 g009
Figure 10. Absorbance of MB solution during photocatalytic degradation using different catalysts under visible light: ZIF-8, ZnFe2O3@ZIF-8 (5 wt%), and ZnFe2O3@ZIF-8 (10 wt%) at 60 min of exposure.
Figure 10. Absorbance of MB solution during photocatalytic degradation using different catalysts under visible light: ZIF-8, ZnFe2O3@ZIF-8 (5 wt%), and ZnFe2O3@ZIF-8 (10 wt%) at 60 min of exposure.
Catalysts 15 00720 g010
Figure 11. Absorbance of MB solution at different visible-light times with (a) ZIF-8 and (b) ZIF-8 (10 wt%) incorporated with ZnFe2O3 catalysts.
Figure 11. Absorbance of MB solution at different visible-light times with (a) ZIF-8 and (b) ZIF-8 (10 wt%) incorporated with ZnFe2O3 catalysts.
Catalysts 15 00720 g011
Figure 12. Concentration decay of MB at different visible-light times with ZIF-8 and ZIF-8 (10 wt%) incorporated with ZnFe2O3 catalysts.
Figure 12. Concentration decay of MB at different visible-light times with ZIF-8 and ZIF-8 (10 wt%) incorporated with ZnFe2O3 catalysts.
Catalysts 15 00720 g012
Figure 13. Photocatalytic efficiency (PCE%) of ZIF-8 and ZIF-8 (10 wt%) incorporated with ZnFe2O3 catalysts in MB.
Figure 13. Photocatalytic efficiency (PCE%) of ZIF-8 and ZIF-8 (10 wt%) incorporated with ZnFe2O3 catalysts in MB.
Catalysts 15 00720 g013
Figure 14. The relation between ln(Ct/Co) and visible-light time of MB in the presence of ZIF-8 and ZIF-8 (10 wt%) incorporated with ZnFe2O3 catalysts.
Figure 14. The relation between ln(Ct/Co) and visible-light time of MB in the presence of ZIF-8 and ZIF-8 (10 wt%) incorporated with ZnFe2O3 catalysts.
Catalysts 15 00720 g014
Figure 15. Recyclability test for ZIF-8 and ZIF-8 (10 wt%) incorporated with ZnFe2O3 catalysts over five successive photocatalytic cycles under visible light.
Figure 15. Recyclability test for ZIF-8 and ZIF-8 (10 wt%) incorporated with ZnFe2O3 catalysts over five successive photocatalytic cycles under visible light.
Catalysts 15 00720 g015
Figure 16. Schematic illustration of the proposed photocatalytic degradation mechanism of methylene blue using ZIF-8@ZnFe2O3 under visible-light irradiation.
Figure 16. Schematic illustration of the proposed photocatalytic degradation mechanism of methylene blue using ZIF-8@ZnFe2O3 under visible-light irradiation.
Catalysts 15 00720 g016
Table 1. Photocatalytic efficiency (PCE%) and apparent rate kinetic constant (kapp) were compared to literature values for MB removal.
Table 1. Photocatalytic efficiency (PCE%) and apparent rate kinetic constant (kapp) were compared to literature values for MB removal.
CatalystLight SourcePCE (%)kₐₚₚ (min−1)Reference
ZIF-8Visible67.87.52 × 10−3Present work
ZIF-8@ZnFe2O Visible9012.51 × 10−3Present work
Fe/ZnO/SiO2 Visible100Not reported[44]
MXeneVisible90.230 × 10−3[45]
NiO-ZnONCsUV721.50 × 10–2 [46]
Cr2O3-CNT NPsVisible689.80 × 10–4[47]
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

Alharbi, F.; Hamdalla, T.; Al-Ghamdi, H.; Albarzan, B.; Darwish, A. Tunable Optical Bandgap and Enhanced Visible Light Photocatalytic Activity of ZnFe2O3-Doped ZIF-8 Composites for Sustainable Environmental Remediation. Catalysts 2025, 15, 720. https://doi.org/10.3390/catal15080720

AMA Style

Alharbi F, Hamdalla T, Al-Ghamdi H, Albarzan B, Darwish A. Tunable Optical Bandgap and Enhanced Visible Light Photocatalytic Activity of ZnFe2O3-Doped ZIF-8 Composites for Sustainable Environmental Remediation. Catalysts. 2025; 15(8):720. https://doi.org/10.3390/catal15080720

Chicago/Turabian Style

Alharbi, Fatma, Taymour Hamdalla, Hanan Al-Ghamdi, Badriah Albarzan, and Ahmed Darwish. 2025. "Tunable Optical Bandgap and Enhanced Visible Light Photocatalytic Activity of ZnFe2O3-Doped ZIF-8 Composites for Sustainable Environmental Remediation" Catalysts 15, no. 8: 720. https://doi.org/10.3390/catal15080720

APA Style

Alharbi, F., Hamdalla, T., Al-Ghamdi, H., Albarzan, B., & Darwish, A. (2025). Tunable Optical Bandgap and Enhanced Visible Light Photocatalytic Activity of ZnFe2O3-Doped ZIF-8 Composites for Sustainable Environmental Remediation. Catalysts, 15(8), 720. https://doi.org/10.3390/catal15080720

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