Next Article in Journal
The Effect of Agitation and the Use of Perfluorodecalin on Lipase Production by Yarrowia lipolytica in a Bioreactor
Previous Article in Journal
Statistical Damage Constitutive Model for Mudstone Based on Triaxial Compression Tests
Previous Article in Special Issue
The Transformative Role of Nano-SiO2 in Polymer Electrolytes for Enhanced Energy Storage Solutions
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Structural, Electrochemical, and Supercapacitor Characterization of Double Metal Oxides Doped Within ZIF-8 Composites

by
Sadeem Saba
1,
Abdulrhman M. Alsharari
1,
Nadiah Y. Aldaleeli
2,
Meshari M. Aljohani
3 and
Taymour A. Hamdalla
1,*
1
Physics Department, Faculty of Science, University of Tabuk, Tabuk 71491, Saudi Arabia
2
Department of Physics, College of Science and Humanities-Jubail, Imam Abdulrahman Bin Faisal University, Jubail 35811, Saudi Arabia
3
Chemistry Department, Faculty of Science, University of Tabuk, Tabuk 71491, Saudi Arabia
*
Author to whom correspondence should be addressed.
Processes 2025, 13(3), 859; https://doi.org/10.3390/pr13030859
Submission received: 12 February 2025 / Revised: 11 March 2025 / Accepted: 13 March 2025 / Published: 14 March 2025
(This article belongs to the Special Issue High-Efficiency Nanomaterials Synthesis and Applications)

Abstract

:
This study investigates the application of double metal oxide zinc ferrite (ZnFe2O3) doped within zeolitic imidazolate framework-8 (ZIF-8) composites for structural, electrochemical, and supercapacitor characterization. The structural characterization has been carried out using XRD, FTIR, DTA, UV-VIS, and HRTEM. By incorporating ZnFe2O3, significant enhancements in the structural integrity and morphology of the ZIF-8 matrices have been achieved, with a decrease in the average crystallite size by about 23%. At 500 °C, the DTA analysis indicated that the weight loss associated with ZnFe2O3 decreased by approximately 5%. The estimated Eg values are 3.08 eV and 3.28 eV for ZIF-8 and ZIF-8@ZnFe2O3, respectively. Regarding the electrochemical performance of the ZIF-8@ZnFe2O3, the anodic peak current density is approximately 0.0025 A (at around 0.5 V), at a scan rate of 10 mV/s. The peak current values increase more rapidly, by about 41%, with increasing scan rate when ZnFe2O3 is present, indicating a synergistic effect between the ZIF-8 and ZnFe2O3 components. The high observed current density peak at 0.03 V can be attributed to the Fe2+/Fe3+ redox couple, facilitated by the ZnFe2O3 component. The ZnFe2O3 addition enhances the electrochemical activity of ZIF-8, leading to increased peak current values at various scan rates. This suggests that the ZnFe2O3 may facilitate charge transfer or improve the conductivity of the material.

1. Introduction

Metal–organic frameworks (MOFs) have emerged as a transformative class of materials due to their unique structural characteristics and multifunctionality [1,2]. Comprising metal nodes coordinated by organic linkers, MOFs exhibit exceptional porosity and tunable chemical environments, making them highly attractive for applications in gas storage, catalysis, and drug delivery [3]. Among these, zeolitic imidazolate frameworks (ZIFs) have gained prominence, particularly ZIF-8, known for its enhanced thermal stability and chemical robustness, which enables it to withstand harsh operating conditions while maintaining a framework structure [4]. The ability to systematically modify the composition and architecture of ZIFs opens new avenues for developing advanced material systems tailored for specific technological needs [5].
Incorporating double metal oxides, such as zinc ferrite (ZnFe2O3), into MOF matrices can significantly advance structural and functional properties [6,7]. ZnFe2O3 is particularly noteworthy due to its favorable electronic and magnetic properties, as well as its capability to absorb light across a broader spectrum, which is critically important for applications in photocatalysis and energy conversion [8]. The strategic doping of ZnFe2O3 into ZIF-8 improves its structural integrity and enhances its optical properties, paving the way for superior light-harvesting capabilities and energy transfer mechanisms [9].
The incorporation of double metal oxides into zeolitic imidazolate framework-8 (ZIF-8) matrices has shown significant promise in enhancing electrochemical and supercapacitor properties. Doping ZIF-8 with double metal oxides, such as zinc ferrite (ZnFe2O3), not only improves the structural stability and conductivity of the composite but also introduces redox-active sites that facilitate charge transfer processes [10]. This synergy enhances the overall electrochemical performance, resulting in higher specific capacitance and improved energy density [11]. The distinctive porosity of ZIF-8 facilitates efficient ion diffusion, while the inclusion of metal oxides enhances the electroactive surface area, optimizing the mechanisms of charge storage. As a result, these doped composites demonstrate exceptional performance in supercapacitor applications, featuring rapid charge and discharge capabilities along with remarkable cycling stability [12]. With targeted modifications, ZnFe2O3-doped ZIF-8 composites can function as advanced and effective energy storage materials, potentially reconciling the need for high energy density with rapid power delivery in electrochemical systems [13].
In this research, the synthesis and characterization of ZnFe2O3-doped ZIF-8 composites have been explored at varying concentrations of dopant. By systematically exploring their structural, optical, and electrochemical properties, we aim to unveil the synergistic effects of this dual-component system, ultimately contributing to the advancement of materials designed for enhanced photonic and energy storage applications. This study not only emphasizes the significance of the incorporated metal oxide but also addresses the potential for innovative applications, extending the functional horizons of MOF-based materials.

2. Experimental Techniques

Advanced experimental equipment was used to measure the structural properties of ZnFe2O3 and the doped ZIF-8 composites. The crystal structure and size were investigated using X-ray diffraction (XRD, D8 Advance X-ray Diffractometer, Bruker, Billerica, MA, USA). The additional functional group, which appeared due to the double metal oxide doping, was investigated using Fourier Transform Infrared Spectroscopy (FTIR, Vertex 70 FTIR Spectrometer, Bruker, Karlsruhe, Germany). The Quantachrome Instrument provided thermal analysis for the prepared sample, while electrochemical analysis was conducted by a potentiostat Ametek, NOVA, Northern Virginia, VA, USA.

3. Sample Fabrication

The chemical used in this research was purchased from Sigma Aldrich. ZnFe2O3-doped ZIF-8 composites can be synthesized using a solvothermal method, which provides a controlled environment for uniform particle formation. The synthesis process was performed as follows: first, zinc nitrate hexahydrate (Zn(NO3)2·6H2O) and 2-methylimidazole (2-MIM) are dissolved in deionized water. The molar ratio of zinc to 2-MIM is maintained at 1:4 to favor ZIF-8 formation. For ZnFe2O3 doping, iron (III) nitrate nonahydrate (Fe(NO3)3·9H2O) is added to achieve the desired weight percentage, specifically 5 wt% of ZnFe2O3 in the final composite. The precursor solution is transferred into a stainless-steel autoclave and heated at 120 °C for 24 h to promote crystallization. After cooling to room temperature, the resultant solid is filtered, washed with deionized water and ethanol, and then dried in an oven at 60 °C for several hours. Finally, the ZnFe2O3-doped ZIF-8 composites are synthesized, and thin films of the material can be developed using a vapor deposition technique. Substrates from glass and quartz wafers are cleaned thoroughly to remove any contaminants that could affect film adhesion and quality. The unit is evacuated to achieve a low-pressure environment, and the material is heated to a temperature that enables sublimation or thermal evaporation of the ZIF-8 powder. As the vapor cools, it condenses onto the substrate, forming a uniform, thin film of ZnFe2O3-doped ZIF-8. The thickness of the film can be controlled by adjusting the deposition time and parameters. Scheme 1 shows the graphical abstract of our experiment.
To evaluate the electrochemical performance of the doped ZIF-8 composites, a glassy carbon electrode (GCE) is prepared. The glassy carbon electrode is polished with alumina slurry (1 µm) to achieve a smooth surface, followed by rinsing with deionized water and drying. In total, 5 mg of the synthesized ZnFe2O3-doped ZIF-8 powder is mixed with a few drops of polyvinyl chloride (PVC) and a small quantity of ethanol in which the slurry is sonicated to ensure uniform dispersion. The prepared GCE is then used for cyclic voltammetry (CV) measurements. The electrochemical behavior of the ZnFe2O3-doped ZIF-8 material can be assessed in suitable electrolyte solutions, allowing for the evaluation of its charge storage capacity and electrochemical kinetics.

4. Results and Discussion

4.1. XRD

The X-ray diffraction (XRD) patterns of ZIF-8 and ZIF-8/ZnFe2O3 are presented in Figure 1. The XRD pattern of pure ZIF-8 shows a series of characteristic peaks at 2θ values of around 7.3°, 12.7°, 32.5°, and 36.8°, which correspond to the (011), (002), (222), and (112) planes, respectively [14]. These peaks are in good agreement with the reported pattern for ZIF-8 [10], indicating the successful synthesis of the material. Upon doping with 5 wt% ZnFe2O3, the XRD pattern of ZIF-8/ZnFe2O3 shows some significant changes. The intensity of the peaks corresponding to the (002) and (112) planes decreases, while the peak corresponding to the (011) plane becomes more pronounced [15]. Additionally, a new peak appears at around 57.5° and 64.3°, which can be attributed to the presence of ZnFe2O3 nanoparticles. The broader peaks in the XRD pattern of ZIF-8/ZnFe2O3 suggest a smaller crystallite size, which may be due to the incorporation of ZnFe2O3 nanoparticles. This is in agreement with previous reports, which have shown that the addition of metal oxide nanoparticles can lead to a decrease in the crystallite size of MOFs [16].
The analysis of the ZIF-8@ZnFe2O3 composite reveals several noteworthy characteristics compared to pure ZIF-8. First, the peaks in the ZIF-8@ZnFe2O3 spectrum are broader, suggesting either a smaller crystallite size or increased disorder within the composite material. Additionally, the peak positions shift slightly to lower 2θ values, indicating a potential expansion of the unit cell; for instance, the (011) peak appears at approximately 7.4° in ZIF-8@ZnFe2O3, compared to 7.6° in ZIF-8. Furthermore, a new peak at around 57.5° in the ZIF-8@ZnFe2O3 spectrum, absent in the ZIF-8 spectrum, may indicate the presence of ZnFe2O3, reflecting some degree of crystallinity in the composite. The average crystallite size of ZIF-8 is estimated to be about 32 nm, while that of ZIF-8@ZnFe2O3 is slightly smaller, around 26 nm. Table 1 shows the XRD analysis of the synthesized ZIF-8/ZnFe2O3 compound. As we can see from the table, the average lattice strain of 0.0031 indicates a relatively low level of distortion within the crystal lattice. This low strain could contribute to maintaining the functionality of the ZIF-8 structure, allowing it to retain its characteristic properties while incorporating the metal oxide. Regarding the lattice parameters, with values of a = 16.33 Å, b = 16.86 Å, and c = 14.91 Å, it appears that there may be some expansion in the unit cell dimensions compared to pure ZIF-8. This expansion might be attributed to the incorporation of ZnFe2O3, which alters the coordination environment around the ZIF-8 framework.

4.2. Thermal Analysis

The Differential Thermal Analysis (DTA) curve in Figure 2 shows the thermal behavior of ZIF-8 and ZIF-8/ZnFe2O3. The DTA curve indicates that the doped sample exhibits improved thermal stability compared to the pristine ZIF-8. The endothermic peak around 500 °C suggests that the doped sample undergoes a phase transition, which might be related to the decomposition of the metal–organic framework. The TGA (Thermogravimetric Analysis) curve (onset Figure 2) shows the weight loss of ZIF-8 and ZIF-8/ZnFe2O3 as a function of temperature. The doped sample exhibits a slower weight loss rate, indicating that the incorporation of ZnFe2O3 enhances the thermal stability of ZIF-8.

4.3. HRTEM

The Transmission Electron Microscopy (TEM) images in Figure 3a,b reveal the morphology and structure of ZIF-8 and ZIF-8/ZnFe2O3. The doped sample shows a more uniform particle size distribution and a higher crystallinity compared to the pristine ZIF-8. The incorporation of ZnFe2O3 might have led to the formation of smaller crystallites, which could enhance the surface area and porosity of the material. The TEM images suggest that the incorporation of ZnFe2O3 affects the crystallite size and morphology of ZIF-8. This could be related to the nucleation and growth mechanisms of ZIF-8 in the presence of ZnFe2O3 nanoparticles. The smaller crystallite size in the doped sample might be beneficial for applications that require high surface areas and porosity, such as catalysis or gas adsorption.

4.4. FTIR

Figure 4 illustrates the Fourier Transform Infrared (FTIR) spectra of ZIF-8 and the composite ZIF-8@ZnFe2O3. The FTIR of ZIF-8 and ZIF-8/ZnFe2O3 exhibit a prominent peak around 3500 cm−1, which is attributed to the stretching vibrations of the (-OH) functional groups. Notably, the peak observed at 1000 cm−1 in ZIF-8 shifts to approximately 900 cm−1 in the composites, indicating alterations in the framework vibrations [16,17]. The spectra exhibit distinct peaks characteristic of ZIF-8, likely linked to framework vibrations and functional groups. Regarding ZIF-8@ZnFe2O3, the first additional peak appears at around 760 cm−1. This peak can be attributed to the Fe-O stretching vibrations in the ZnFe2O3 nanoparticles. Incorporating ZnFe2O3 into ZIF-8 leads to the formation of metal–oxygen bonds, resulting in this new peak. The second additional peak appears at around 1630 cm−1, which can be assigned to the C=O stretching vibrations of the 2-methylimidazole linkers in ZIF-8. Finally, a peak around 3500 cm−1 may occur because the water molecules can give rise to O-H stretching vibrations [18,19]. The incorporation of ZnFe2O3 could improve charge transfer and increase the stability of the composite, making it a promising candidate for advanced materials in energy conversion and storage technologies.

4.5. Energy Gap

The energy gap calculation for ZIF-8 and ZIF-8@ZnFe2O3 are introduced in Figure 5. This figure depicts the Tauc’s plots used to estimate the energy gap (Eg) of the materials. As we can see from the figure, the calculated Eg values are 3.08 eV for ZIF-8 and 3.28 eV for the ZIF-8@ZnFe2O3 composite. For ZIF-8, the peaks in the UV region can be attributed to the electronic transitions associated with the metal–organic framework’s (MOF) ligands, while additional peaks noted in the ZIF-8@ZnFe2O3 composite may result from the presence of ZnFe2O3, leading to new energy levels within the bandgap. This incorporation can create defect states, which may contribute to enhanced light absorption and influence carrier dynamics in the material. The observed increase in the energy gap for the composite indicates that the electronic interactions between ZIF-8 and ZnFe2O3 modify the electronic structure, potentially leading to improved stability and performance in energy-related applications [20,21]. Indirect transitions can occur when the electron must transition through an intermediate state, leading to a different optical behavior than direct transitions. Previous studies carried out by our team and others have documented similar behaviors in analogous materials [10,16,22]. Research has shown that metal–organic frameworks and their composites frequently display indirect band gaps due to their unique structural and electronic configurations. These findings align with my observation of the indirect absorption mechanism within our studied materials. The absorption spectra for ZIF-8 and ZIF-8@ZnFe2O3 suggest a tailing at higher energies, indicating the presence of indirect transitions. Such behavior often necessitates analyzing the data with an exponent of 1/2 to accurately describe the optical properties.

4.6. CV Measurements

The CV curves in Figure 6 show the charging and discharging behavior of ZIF-8 and ZIF-8/ZnFe2O3 at different scan rates (10 mV/s, 30 mV/s, and 50 mV/s). The doped sample exhibits a more rectangular shape, indicating a higher electrochemical stability and capacitive behavior. The CV curves in Figure 6 suggest that the doped sample undergoes redox reactions during the charging and discharging process. The presence of ZnFe2O3 might promote the formation of redox-active species, which could contribute to improved electrochemical performance. As you can see from the figure, the peak current density for the oxidation reaction (anodic peak) is higher for ZIF-8@ZnFe2O3 compared to ZIF-8. For ZIF-8, the anodic peak current density is approximately 0.0012 A (at around 0.5 V) at a scan rate of 10 mV/s. In contrast, for ZIF-8@ZnFe2O3, the anodic peak current density is approximately 0.0025 A (at around 0.5 V) at a scan rate of 10 mV/s. This represents a 108% increase in the anodic peak current density. Similarly, the cathodic peak current density is also higher for ZIF-8@ZnFe2O3. For ZIF-8, the cathodic peak current density is approximately −0.0010 A (at around 0.2 V) at a scan rate of 10 mV/s. In contrast, for ZIF-8@ZnFe2O3, the cathodic peak current density is approximately −0.0018 A (at around 0.2 V) at a scan rate of 10 mV/s. This represents an 80% increase in the cathodic peak current density. These values provide evidence that the introduction of ZnFe2O3 has improved the charging and discharging capabilities of the material, as demonstrated by the increased peak current densities.
The peak current density of ZIF-8 and ZIF-8/ZnFe2O3 is shown in Figure 7. The doped sample exhibits a higher peak current density, suggesting an improvement in electrochemical activity. The peak current values seem to increase more rapidly, by about 41%, with increasing scan rate when Fe2O3 is present, indicating a synergistic effect between the ZIF-8 and Fe2O3 components. The Fe2O3 addition enhances the electrochemical activity of ZIF-8, leading to increased peak current values at various scan rates. This suggests that the Fe2O3 may be facilitating charge transfer or improving the conductivity of the material.
The current density of ZIF-8/ZnFe2O3 as a function of potential is shown in Figure 8. The doped sample exhibits a higher current density, indicating a higher electrochemical activity. The curve exhibits characteristic redox behavior, with a distinct peak observed at approximately 0.03 V, indicating the onset of electrochemical activity. Notably, the peak current density reaches a value of around 0.6 mA/cm2, signifying a significant enhancement in electrochemical performance compared to pristine ZIF-8. The incorporation of ZnFe2O3 into the ZIF-8 framework has resulted in a substantial increase in current density, particularly in the positive potential region. This implies that the modification has led to an improvement in the material’s conductivity and electrochemical reactivity. The high observed peak at 0.03 V can be attributed to the Fe2+/Fe3+ redox couple, which is facilitated by the ZnFe2O3 component. The emergence of this peak suggests that the ZnFe2O3 has effectively introduced electroactive sites within the ZIF-8 framework, thereby enhancing the material’s electrochemical performance.
Figure 9 shows the specific capacitance of ZIF-8/ZnFe2O3 as a function of voltage rate. The doped sample exhibits a higher specific capacitance, which could be related to ZnFe2O3 nanoparticles acting as a conductive pathway, enhancing the electrochemical activity of ZIF-8. The incorporation of ZnFe2O3 might increase the surface area and porosity of ZIF-8, allowing for better electrolyte diffusion and ion transport [17]. The ZnFe2O3 nanoparticles might promote the formation of redox-active species, contributing to improved electrochemical performance [18,23].
Figure 10 presents the galvanostatic charge–discharge (GCD) characteristics for ZIF-8/glassy carbon electrode (GCE) and ZIF-8/ZnFe2O3/GCE composites. The charge and discharge curves for both materials demonstrate a high degree of symmetry and exhibit nearly triangular shapes across various current densities. This symmetry is indicative of the ideal electrochemical behavior associated with capacitive materials, reflecting efficient charge storage and transfer mechanisms. The triangular profile suggests that the composites maintain their capacitive performance even at elevated current densities, which is crucial for practical applications in energy storage devices. These results underscore the commendable capacitive properties of the materials, highlighting their potential for usage in supercapacitors where rapid charge and discharge cycles are essential.
A cycling stability test was performed over 2000 cycles to assess the durability of the ZIF-8/GCE and ZIF-8/ZnFe2O3/GCE, as illustrated in Figure 11. The results indicate that after 2000 cycles, the ZIF-8/ZnFe2O3/GCE retained approximately 95% of its original capacity, demonstrating impressive stability throughout the cyclic voltammetry (CV) testing. This high retention of capacity reaffirms that the electrode maintains its performance over extended use, which is crucial for practical applications in energy storage devices. Furthermore, the improved electrochemical performance of the ZIF-8/ZnFe2O3/GCE can be attributed to the alterations observed in the complex impedance, represented as (Z* = Z′ + jZ″). This change in impedance suggests enhanced conductivity and charge transfer capabilities within the composite, contributing to its superior stability and efficiency during cycling. Overall, these findings highlight the promising potential of the ZIF-8/ZnFe2O3 composite for use in long-lasting supercapacitor applications.
Figure 12 illustrates the relationship between the real part of the impedance (Z′) and the imaginary part of the impedance (Z″) for the ZIF-8/GCE and the ZIF-8/ZnFe2O3/GCE composites. From the figure, it is evident that the incorporation of ZnFe2O3 results in an increase in bulk resistance, rising from 1400 kΩ for the pure ZIF-8/GCE to 1800 kΩ for the ZIF-8/ZnFe2O3/GCE. When combined with ZnFe2O, the overall electrical performance of the composite is adversely affected. This reduction in conductivity leads to an increase in the bulk resistance observed in the ZIF-8/ZnFe2O3/GCE compared to the pure activated ZIF-8. The increased bulk resistance may hinder charge transport within the composite, potentially affecting its electrochemical performance. Therefore, while the addition of ZnFe2O3 can enhance specific properties, such as electrochemical activity, the overall conductivity balance illustrated in Figure 12 underscores the importance of optimizing composite formulations to achieve desirable performance in electrochemical applications.
In the zinc ferrite (ZnFe2O3)-doped zeolitic imidazolate framework-8 (ZIF-8) composites, the charge storage mechanism is enhanced significantly through a dual action: Faradaic redox reactions and electrical double-layer capacitance. Table 2 shows a comparison between electrochemical performance for ZIF-based composites. The studied composite shows a notable increase in peak current density; for instance, the anodic peak current density observed is approximately 0.0025 A at around 0.5 V with a scan rate of 10 mV/s. This value increases by about 41% with rising scan rates when ZnFe2O3 is present, illustrating the synergy between ZIF-8 and ZnFe2O3 in facilitating charge transfer. Structural characterization indicates a decrease in average crystallite size of about 23% due to the incorporation of ZnFe2O3, which enhances the material’s surface area and active sites for charge storage. Furthermore, the estimated electrochemical bandgap values of ZIF-8 and ZIF-8@ZnFe2O3 are 3.08 eV and 3.28 eV, respectively, suggesting improved electronic conductivity in the composite. The presence of the Fe2+/Fe3+ redox couple is crucial, as it can lead to significant capacitive contributions, enabling sustained performance with higher current densities (as evidenced by the observed peak at 0.03 V) and enhancing the overall efficiency of energy storage, making these composites promising candidates for advanced supercapacitor applications.

5. Conclusions

The synthesis and characterization of ZnFe2O3-doped ZIF-8 composites have been carried out. Through a systematic examination of their structural, optical, and electrochemical properties, we aspire to reveal the synergistic effects of this dual-component system, thereby contributing to the development of materials intended for improved photonic and energy storage applications. The doped sample exhibits a slower weight loss rate, indicating that the incorporation of ZnFe2O3 enhances the thermal stability of ZIF-8. The TEM image for the ZIF-8@ZnFe2O3 doped sample shows a more uniform particle size distribution and a higher crystallinity compared to the pristine ZIF-8. The CV curves for the doped sample show a more rectangular shape than pristine ZIF-8, indicating improved electrochemical stability and capacitive behavior. The peak current densities increased by 108% for the anodic peak and 80% for the cathodic peak, demonstrating better charge storage and transfer capabilities of the composite. The addition of ZnFe2O3 enhances electron exchange and helps form redox-active species during electrochemical processes. The GCD results support these findings, displaying symmetrical charge and discharge curves across various current densities. ZnFe2O3 not only boosts the electrochemical activity of ZIF-8 by providing more active sites but also improves the material’s overall conductivity, which is vital for energy storage. The redox peaks related to Fe2+/Fe3+ transitions further confirm the significant enhancements achieved with this composite. Overall, these results highlight the ZIF-8/ZnFe2O3 composite’s strong potential for advanced energy storage applications, thanks to its rapid charge–discharge capabilities and excellent electrochemical stability, making it suitable for supercapacitors and similar technologies.

Author Contributions

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

Funding

This work was funded by the Deanship of Research and Graduate Studies at the University of Tabuk through Research No. 0018-2024-S.

Data Availability Statement

The authors confirm that the datasets used and analyzed during the current study are available from the corresponding author on reasonable request.

Acknowledgments

The authors extend their appreciation to the Deanship of Research and Graduate Studies at the University of Tabuk for funding this work through Research no. 0018-2024-S.

Conflicts of Interest

The authors declare that there is no conflict of interest in the manuscript.

References

  1. Li, B.; Wen, H.M.; Cui, Y.; Zhou, W.; Qian, G.; Chen, B. Emerging multifunctional metal–organic framework materials. Adv. Mater. 2016, 28, 8819–8860. [Google Scholar] [CrossRef] [PubMed]
  2. Jiao, L.; Seow, J.Y.R.; 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]
  3. Liu, J.; Huang, J.; Zhang, L.; Lei, J. Multifunctional metal–organic framework heterostructures for enhanced cancer therapy. Chem. Soc. Rev. 2021, 50, 1188–1218. [Google Scholar] [CrossRef]
  4. Alfadhli, S.; Darwish, A.A.A.; Al-Ghamdi, S.A.; Aljohani, M.M.; Seleim, S.M.; Mahmoud, M.E.; Khasim, S.; Hamdalla, T.A. Synthesis, characterization, optical, and sensing investigations for Fe-BDC doped with 10 wt.% of activated food waste biochar. J. Asian Ceram. Soc. 2024, 12, 23–33. [Google Scholar] [CrossRef]
  5. Alsharif, M.A.; Alqurashi, R.S.; Hamdalla, T.A. The role of annealing in optimizing the electrical, linear, and nonlinear optical properties ZIf-8@ TiO2 films for possible use in photonic and optoelectronic devices. Phys. Scr. 2024, 99, 065983. [Google Scholar] [CrossRef]
  6. Li, Y.; Xu, Y.; Yang, W.; Shen, W.; Xue, H.; Pang, H. MOF-derived metal oxide composites for advanced electrochemical energy storage. Small 2018, 14, 1704435. [Google Scholar] [CrossRef]
  7. Das, R.; Pachfule, P.; Banerjee, R.; Poddar, P. Metal and metal oxide nanoparticle synthesis from metal organic frameworks (MOFs): Finding the border of metal and metal oxides. Nanoscale 2012, 4, 591–599. [Google Scholar] [CrossRef]
  8. Wang, Y.C.; Li, W.B.; Zhao, L.; Xu, B.Q. MOF-derived binary mixed metal/metal oxide@ carbon nanoporous materials and their novel supercapacitive performances. Phys. Chem. Chem. Phys. 2016, 18, 17941–17948. [Google Scholar] [CrossRef]
  9. Chen, J.J.; Chen, Y.T.; Senthil Raja, D.; Kang, Y.H.; Tseng, P.C.; Lin, C.H. Metal-organic frameworks to metal/metal oxide embedded carbon matrix: Synthesis, characterization and gas sorption properties. Materials 2015, 8, 5336–5347. [Google Scholar] [CrossRef]
  10. Hamdalla, T.A.; Al-Ghamdi, S.A.; Alfadhli, S.; Alsharari, A.M.; Chiesa, M.; Khasim, S. PEDOT: PSS Doped Activated Biochar as a Novel Composite Material for Photocatalytic and Efficient Energy Storage Application. Catalysts 2024, 14, 630. [Google Scholar] [CrossRef]
  11. An, C.; Zhang, Y.; Guo, H.; Wang, Y. Metal oxide-based supercapacitors: Progress and prospectives. Nanoscale Adv. 2019, 1, 4644–4658. [Google Scholar] [CrossRef] [PubMed]
  12. Nguyen, T.; Montemor, M.d.F. Metal oxide and hydroxide–based aqueous supercapacitors: From charge storage mechanisms and functional electrode engineering to need-tailored devices. Adv. Sci. 2019, 6, 1801797. [Google Scholar] [CrossRef]
  13. Yuan, S.; Duan, X.; Liu, J.; Ye, Y.; Lv, F.; Liu, T.; Wang, Q.; Zhang, X. Recent progress on transition metal oxides as advanced materials for energy conversion and storage. Energy Storage Mater. 2021, 42, 317–369. [Google Scholar] [CrossRef]
  14. Khan, I.U.; Othman, M.H.D.; Ismail, A.F.; Ismail, N.; Jaafar, J.; Hashim, H. Structural transition from two-dimensional ZIF-L to three-dimensional ZIF-8 nanoparticles in aqueous room temperature synthesis with improved CO2 adsorption. Mater. Charact. 2018, 136, 407–416. [Google Scholar] [CrossRef]
  15. Yin, R.; Liu, M.; Tang, R.; Yin, L. CdS Nanoparticle-Modified α-Fe2O3/TiO2 Nanorod Array Photoanode for Efficient Photoelectrochemical Water Oxidation. Nanoscale Res. Lett. 2017, 12, 1–8. [Google Scholar] [CrossRef] [PubMed]
  16. Al-atawi, S.; Aljohani, M.M.; Hamdalla, T.A.; Al-Ghamdi, S.A.; Alsharari, A.M.; Khasim, S. Novel Biochar-Modified ZIF-8 Metal–Organic Frameworks as a Potential Material for Optoelectronic and Electrochemical Energy Storage Applications. Catalysts 2024, 14, 705. [Google Scholar] [CrossRef]
  17. Zhang, Z.; Xian, S.; Xi, H.; Wang, H.; Li, Z. Improvement of CO2 adsorption on ZIF-8 crystals modified by enhancing basicity of surface. Chem. Eng. Sci. 2011, 66, 4878–4888. [Google Scholar] [CrossRef]
  18. Yu, K.; Lou, L.L.; Liu, S.; Zhou, W. Asymmetric oxygen vacancies: The intrinsic redox active sites in metal oxide catalysts. Adv. Sci. 2020, 7, 1901970. [Google Scholar] [CrossRef]
  19. Manikandan, V. Efficient Photocatalytic Degradation of Crystal Violet and Rhodamine B Dyes Using RGO@ ZnFe2O3 Nanocomposite. Strength Mater. 2023, 55, 167–178. [Google Scholar] [CrossRef]
  20. Guan, B.Y.; Yu, X.Y.; Wu, H.B.; Lou, X.W. Complex nanostructures from materials based on metal–organic frameworks for electrochemical energy storage and conversion. Adv. Mater. 2017, 29, 1703614. [Google Scholar] [CrossRef]
  21. Zhou, W.; Tang, Y.; Zhang, X.; Zhang, S.; Xue, H.; Pang, H. MOF derived metal oxide composites and their applications in energy storage. Coord. Chem. Rev. 2023, 477, 214949. [Google Scholar] [CrossRef]
  22. Cardoso, J.C.; Stulp, S.; De Brito, J.F.; Flor, J.B.S.; Frem, R.C.G.; Zanoni, M.V.B. MOFs based on ZIF-8 deposited on TiO2 nanotubes increase the surface adsorption of CO2 and its photoelectrocatalytic reduction to alcohols in aqueous media. Appl. Catal. B Environ. 2018, 225, 563–573. [Google Scholar] [CrossRef]
  23. Wang, Z.; Pakoulev, A.; Pang, Y.; Dlott, D.D. Vibrational substructure in the OH stretching band of water. Chem. Phys. Lett. 2003, 378, 281–288. [Google Scholar] [CrossRef]
  24. Sadem, A.; Aljohani, M.M.; Al-Ghamdi, S.A.; Alsharari, A.M.; Sadque, M.; Hamdalla, T.A. The Optimal Doping Ratio of Fe2O3 for Enhancing the Electrochemical Stability of Zeolitic Imidazolate Framework-8 for Energy Storage Devices. Adv. Condens. Matter Phys. 2024, 2024, 5134666. [Google Scholar]
  25. Zhang, P.; Sun, F.; Shen, Z.; Cao, D. ZIF-derived porous carbon: A promising supercapacitor electrode material. J. Mater. Chem. A 2014, 2, 12873–12880. [Google Scholar] [CrossRef]
  26. Li, X.; Hao, C.; Tang, B.; Wang, Y.; Liu, M.; Wang, Y.; Zhu, Y.; Lu, C.; Tang, Z. Supercapacitor electrode materials with hierarchically structured pores from carbonization of MWCNTs and ZIF-8 composites. Nanoscale 2017, 9, 2178–2187. [Google Scholar] [CrossRef]
Scheme 1. The graphical abstract of the experimental procedure.
Scheme 1. The graphical abstract of the experimental procedure.
Processes 13 00859 sch001
Figure 1. XRD patterns of (a) ZIF-8 and (b) ZIF-8/ZnFe2O3.
Figure 1. XRD patterns of (a) ZIF-8 and (b) ZIF-8/ZnFe2O3.
Processes 13 00859 g001
Figure 2. DTA of ZIF-8 and ZIF-8/ZnFe2O3, with onset graph depicting TGA results.
Figure 2. DTA of ZIF-8 and ZIF-8/ZnFe2O3, with onset graph depicting TGA results.
Processes 13 00859 g002
Figure 3. (a) TEM images of ZIF-8. (b) TEM images of ZIF-8/ZnFe2O3.
Figure 3. (a) TEM images of ZIF-8. (b) TEM images of ZIF-8/ZnFe2O3.
Processes 13 00859 g003
Figure 4. FTIR of (a) ZIF-8 and (b) ZIF-8/ZnFe2O3.
Figure 4. FTIR of (a) ZIF-8 and (b) ZIF-8/ZnFe2O3.
Processes 13 00859 g004
Figure 5. Energy gap calculation for ZIF-8 and ZIF-8@ZnFe2O3.
Figure 5. Energy gap calculation for ZIF-8 and ZIF-8@ZnFe2O3.
Processes 13 00859 g005
Figure 6. Charging and discharging behavior of ZIF-8 and ZIF-8/ZnFe2O3.
Figure 6. Charging and discharging behavior of ZIF-8 and ZIF-8/ZnFe2O3.
Processes 13 00859 g006
Figure 7. Peak current for ZIF-8 and ZIF-8/ZnFe2O3.
Figure 7. Peak current for ZIF-8 and ZIF-8/ZnFe2O3.
Processes 13 00859 g007
Figure 8. Current density for ZIF-8/ZnFe2O3.
Figure 8. Current density for ZIF-8/ZnFe2O3.
Processes 13 00859 g008
Figure 9. Specific capacitance for ZIF-8/ZnFe2O3.
Figure 9. Specific capacitance for ZIF-8/ZnFe2O3.
Processes 13 00859 g009
Figure 10. The galvanostatic charge–discharge (GCD) characteristics for ZIF-8/glassy carbon electrode (GCE) and ZIF-8/ZnFe2O3/GCE composites.
Figure 10. The galvanostatic charge–discharge (GCD) characteristics for ZIF-8/glassy carbon electrode (GCE) and ZIF-8/ZnFe2O3/GCE composites.
Processes 13 00859 g010
Figure 11. Cycling stability over 2000 cycles to assess the durability of the ZIF-8/GCE and ZIF-8/ZnFe2O3/GCE.
Figure 11. Cycling stability over 2000 cycles to assess the durability of the ZIF-8/GCE and ZIF-8/ZnFe2O3/GCE.
Processes 13 00859 g011
Figure 12. The real part of the impedance (Z′) and the imaginary part of the impedance (Z″) for the ZIF-8/GCE and the ZIF-8/ZnFe2O3/GCE composites.
Figure 12. The real part of the impedance (Z′) and the imaginary part of the impedance (Z″) for the ZIF-8/GCE and the ZIF-8/ZnFe2O3/GCE composites.
Processes 13 00859 g012
Table 1. XRD analysis of the synthesized ZIF-8/ZnFe2O3 compound.
Table 1. XRD analysis of the synthesized ZIF-8/ZnFe2O3 compound.
No.Compound2θ°D-Spacing [Å]a, b, ch, k, l(Crystallite Size) D, (nm)Lattice Strain
(nm)
1ZIF-8/ZnFe2O312.7°3.70a = 8.33
b = 7.86
c = 6.91
(002)17.0940.00142
232.5°3.66(222)32.0650.00374
336.8°3.95(112)29.4110.00264
457.5°2.83(440)21.5210.00612
564.3°3.91(442)41.3410.00180
Table 2. Comparison between electrochemical performance for ZIF-based composites.
Table 2. Comparison between electrochemical performance for ZIF-based composites.
ZIF-8 CompositeSpecific
Capacitance F/G
Current Density Peak mA/m2Ref.
ZIF-8/ZnFe2O33853.5Present work
ZIF-8@RHBC4523.7[24]
ZIF-7/glucose2282.9[25]
MWCNT/ZIF-8293.43.5[26]
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

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. https://doi.org/10.3390/pr13030859

AMA Style

Saba S, Alsharari AM, Aldaleeli NY, Aljohani MM, Hamdalla TA. Structural, Electrochemical, and Supercapacitor Characterization of Double Metal Oxides Doped Within ZIF-8 Composites. Processes. 2025; 13(3):859. https://doi.org/10.3390/pr13030859

Chicago/Turabian Style

Saba, Sadeem, Abdulrhman M. Alsharari, Nadiah Y. Aldaleeli, Meshari M. Aljohani, and Taymour A. Hamdalla. 2025. "Structural, Electrochemical, and Supercapacitor Characterization of Double Metal Oxides Doped Within ZIF-8 Composites" Processes 13, no. 3: 859. https://doi.org/10.3390/pr13030859

APA Style

Saba, S., Alsharari, A. M., Aldaleeli, N. Y., Aljohani, M. M., & Hamdalla, T. A. (2025). Structural, Electrochemical, and Supercapacitor Characterization of Double Metal Oxides Doped Within ZIF-8 Composites. Processes, 13(3), 859. https://doi.org/10.3390/pr13030859

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