Previous Article in Journal
The EuAPS Betatron Radiation Source: Status Update and Photon Science Perspectives
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Enhancing the Photoelectrochemical Performance of a Superlattice p–n Heterojunction CuFe2O4/ZnFe2O4 Electrode for Hydrogen Production

by
M. K. Al Turkestani
Department of Physics, Umm Al-Qura University, Makkah 24382, Saudi Arabia
Condens. Matter 2024, 9(3), 31; https://doi.org/10.3390/condmat9030031
Submission received: 3 July 2024 / Revised: 24 July 2024 / Accepted: 25 July 2024 / Published: 26 July 2024
(This article belongs to the Section Physics of Materials)

Abstract

:
A p–n heterojunction film consisting of p-type CuFe2O4 and n-type ZnFe2O4 was fabricated in this study. The n-type ZnFe2O4 film was deposited on a stainless steel substrate using the spray pyrolysis method, after which a top layer of p-type CuFe2O4 thin film was deposited and annealed. Characterization techniques, such as X-ray diffraction, scanning electron microscopy, UV–Vis diffuse reflectance spectroscopy, and photoluminescence, confirmed the formation of a superlattice p–n heterojunction between CuFe2O4 and ZnFe2O4. Photoelectrochemical measurements were conducted to investigate the photoelectrochemical properties of the samples, resulting in a photocurrent of 1.2 mA/cm2 at 1.5 V (vs. Ag/AgCl) under illumination from a 100-watt LED light source. Utilizing the p–n junction of CuFe2O4/ZnFe2O4 as a photoanode increased the hydrogen production rate by 30% compared to that of the dark measurement. This enhancement in performance was attributed to the potential barrier at the p–n heterojunction interface, which improved the separation of photoinduced electron–hole pairs and facilitated a more efficient charge transfer. Additionally, coating the stainless steel electrode with this ferrite sample improved both the corrosion resistance and the stability of hydrogen production over extended operation times.

1. Introduction

Producing hydrogen as an alternative source of clean energy has become a significant challenge [1,2]. Photoelectrochemical cells (PECs) have been extensively studied since it was discovered that TiO2 can split water under UV irradiation [3,4,5,6]. However, these PEC cells have limited efficiency due to the use of UV radiation, which accounts for only 4% of the total solar irradiation spectrum. To improve hydrogen production under solar light, which contains a large portion (43%) of visible radiation [7,8], there is a need to develop a new photocatalyst with a smaller energy gap in the visible range, chemical stability, low electron–hole recombination rate, and nontoxic properties.
Magnetic spinel ferrites (MFe2O4; M = Cu, Ni, Cd, Ca, Zn, etc.) have emerged as promising materials for hydrogen production [9,10]. These ferrites exhibit high thermal and chemical stability [11,12,13] and possess a narrow energy gap in the visible range [14,15,16], making them suitable for photocatalysis and applications involving solar water splitting [17,18,19,20]. Combining p-type ferrites with n-type semiconductors to form p–n heterojunctions, such as CuFe2O4/CdS, CuFe2O4/SnO2, CuFe2O4/TiO2, CaFe2O4/WO3, CaFe2O4/BiVO4, and CsFe2O4/Fe2O3, has improved PEC performance compared to bare n-type materials [21,22,23,24,25,26,27]. The enhanced photocatalytic activity is attributed to the separation of photogenerated electron–hole pairs by the internal potential barrier at the p–n heterojunction interface. However, the challenge lies in the chemical and physical structure mismatch between the p-type and n-type materials, resulting in low PEC cell efficiency due to defects in the interface.
In this paper, a superlattice heterojunction between p-type CuFe2O4 and n-type ZnFe2O4 is investigated as a photoanode for hydrogen production. Using these ferrites offers several advantages: (1) they have the same chemical and crystal structure, (2) their lattice mismatch is minimal, (3) their energy gaps are in the visible range, and (4) they are expected to generate a high internal potential at the p–n junction interface.
Throughout this study, we employed various techniques to characterize the structural and optical properties of the samples. Additionally, linear sweep voltammetry (LSV) was used to evaluate the performance of the PEC cell, and the hydrogen production rate was monitored over an extended period of operation time.

2. Sample Preparation

2.1. Preparation of CuFe2O4 and ZnFe2O4 Films on Stainless Steel Sheets

CuFe2O4 and ZnFe2O4 films were fabricated on stainless steel 304 plates using the spray pyrolysis technique. Initially, the stainless steel sheets were cut into 5 cm × 5 cm pieces. To increase the surface area, the plates were etched with 1200 grinding papers for 5 min. Then, the plates underwent ultrasonic cleaning in ethanol and deionized water for 10 min each. Finally, the stainless steel plates were dried using a hot air path to prepare them for the next steps. For depositing the desired metal ferrites, Fe(NO3)3·7H2O, Cu(NO3)2·5H2O, and Zn(NO3)2·6H2O salts from Merck India with 98% purity were utilized as sources of Fe, Cu, and Zn ions, respectively. All chemicals were from the molar ratios of Fe:Zn and Fe:Cu, which were maintained at 2:1. The solvent used in this study was a mixture of water and ethanol with a volume ratio of 3:1. The solution was sprayed on stainless steel plates using an ultrasonic nebulizer at a temperature of 350 °C. Finally, the deposited ferrite films were annealed using air heated to 500 °C for 2 h. The thickness of the prepared films was controlled by the deposition time. The film thickness was estimated using the weight difference method, revealing that a deposition time of 30 min resulted in films with a thickness of 2 μm.

2.2. Preparation of CuFe2O4/ZnFe2O4 P–N Junction

One of the annealed Zn ferrite films with a thickness of 2 μm served as the substrate for depositing the Cu ferrite film using the same steps, except that the deposition time was reduced to 15 min, and the annealing temperature was set to 450 °C for 1 h. The thickness of the Cu ferrite film was approximately 200 nm. Confirmation of the CuFe2O4/ZnFe2O4 p–n junction formation was achieved by measuring the electric resistance in forward bias (indicating low resistance) and reverse bias (indicating high resistance).

2.3. Characterization Techniques

In this study, the crystal structure and phase identification were investigated using the X-ray diffractometer Philips model (PW-1710, Philips, Amsterdam, The Netherlands) with Cu-Kα radiation (λ = 1.5406 Å). The surface morphologies of the thin films were examined using field emission scanning electron microscopy. To determine the bandgap energy value, the thin films were examined using a UV–Vis model JASCO 670 spectrophotometer (Jasco, Tokyo, Japan) within the wavelength range of 200–800 nm. The electrochemical performances of the deposited films were assessed in a standard three-electrode cell at room temperature. The anode electrode was prepared for each of the thin films, a graphene electrode served as the counter electrode, and an Ag/AgCl reference electrode was employed.
Electrochemical impedance spectroscopy (EIS) and Mott–Schottky plots were carried out using a TEGAM LRC Meter Model 3550 (TEGAM, Geneva, OH, USA). Cyclic voltammetry (CV) studies of the samples were conducted in a 0.3 M NaOH aqueous electrolyte using a specially designed linear sweep voltammetry (LSV) device. This LSV device was calibrated using an electrochemical workstation (Potentiostat/Galvanostat PARSTAT 2273 model, Princeton Applied Research, Oak Ridge, TN, USA). The hydrogen production rate was measured by connecting the generator to an inverted measuring cylinder for a specific period to quantify the production rate.

3. Results and Discussion

3.1. X-ray Diffraction Measurements

Figure 1 presents the X-ray diffraction (XRD) patterns obtained for the prepared samples. The XRD patterns of the bare stainless steel exhibit characteristic peaks corresponding to the stainless steel’s 304 phase. Specifically, the peaks observed at 34°, 51°, and 74° can be attributed to the face-centered cubic (FCC) structure of stainless steel 304 [28]. For the ferrite samples, the XRD patterns indicate the formation of a single-spinel cubic-phase structure on the stainless steel substrate. By applying Bragg’s law and the Williamson–Hall equation to the three most intense peaks, various parameters, such as the lattice parameter, crystallite size, and microstrain, were calculated and are listed in Table 1.
It is evident that the lattice parameter of Zn ferrite (aZn = 8.425 Å) is greater than that of Cu ferrite (aCu = 8.384 Å) [29,30]. This difference can be attributed to the larger ionic radius of Zn2+ (0.78 nm) compared to that of Cu2+ (0.65 nm). Furthermore, the crystallite size of Cu ferrite is larger than that of Zn ferrite, indicating a higher degree of crystallinity in Cu ferrite.
Additionally, the lattice parameter of the CuFe2O4/ZnFe2O4 sample was between aZn and aCu but leaned toward aZn. However, the microstrain showed a significant increase. These results confirm the presence of a superlattice structure consisting of Cu ferrite film on Zn ferrite. The increased microstrain can be attributed to the lattice mismatch (∆a/aZn = (aZn − aCu)/aZn = 1%) between Cu ferrite and Zn ferrite. Finally, the crystallite size of this sample is greater than that of both Cu ferrite and Zn ferrite, indicating an enhancement in the crystallinity of the Cu ferrite film when grown on Zn ferrite, with the Zn ferrite crystals acting as seeds for the growth of the Cu ferrite film.

3.2. Optical Absorption and Photoluminescence

The diffuse reflection method was employed to determine the optical bandgap of the prepared samples. Figure 2 illustrates the reflection coefficient (R) of all the investigated samples in the UV–Vis range.
The absorption coefficient (α) can be estimated using the following Kubelka–Munk equation [31]:
α = (1 − R)2/2R
The direct energy gap of the prepared sample can be determined using Tauc’s relation [32], expressed as follows:
(αhν)2 = A(hν − Eg)
Here, A represents a constant, hν denotes the photon energy, and Eg represents the optical direct energy gap.
Figure 3 displays Tauc’s relation between (αhν)2 and hν. The energy gap can be determined by identifying the intersections of the lines with hν axis. It was observed that the energy gap of Zn ferrite was 2.49 eV, while that of Cu ferrite was 2.04 eV (inset figure). These results are consistent with previous reports [33,34,35]. For the CuFe2O4/ZnFe2O4 sample, it is crucial to note the presence of two transition states corresponding to Cu ferrite and Zn ferrite. It is well established that CuFe2O4 is a p-type semiconductor with a smaller energy gap than ZnFe2O4, which is an n-type semiconductor. Consequently, a space charge region and a potential barrier formed at the interface between the two ferrites to achieve Fermi-level continuity, and the energy gap for CuFe2O4 on the ZnFe2O4 layer decreased to 1.78 eV. This decrease could be a result of the increase in the grain size of CuFe2O4, as it was grown on ZnFe2O4 [36].
The presence of this barrier potential at the interface can reduce the recombination rate between electrons and holes, thereby influencing the photoluminescent intensity. Therefore, to determine the optical properties, photoluminescence (PL) measurements were conducted.
Figure 4 displays the PL emission spectra obtained at an excitation wavelength of 325 nm (Eexc. = 3.8 eV), which exceeds the energy gaps of Zn ferrite and Cu ferrite. It is evident that the PL signal is stronger for Cu ferrite than for Zn ferrite, indicating a faster recombination rate of electron–hole pairs in Cu ferrite. Conversely, a slight decrease in the PL signal was observed for the p–n junction sample. This result can be attributed to the presence of a strong barrier potential at the CuFe2O4/ZnFe2O4 interface, as mentioned earlier.

3.3. Electrochemical Impedance Spectroscopy

The basic photoelectrochemical cell (PEC) parameters of fabricated photoelectrodes, including CuFe2O4, ZnFe2O4, and CuFe2O4/ZnFe2O4, were investigated under dark conditions using electrochemical impedance spectroscopy (EIS). EIS measurements were conducted with a voltage amplitude of 50 mV over the frequency range of 1.0 Hz–100 kHz. The resulting Nyquist plots (Figure 5) display distinct semicircles, indicating the charge transfer process at the electrode–electrolyte interface, while the semicircle diameter represents the charge transfer resistance.
Notably, the CuFe2O4 sample exhibited the largest diameter, indicative of slow charge transfer at the electrode–electrolyte interface, whereas the ZnFe2O4 sample displayed a smaller radius, signifying rapid and effective charge transfer [36,37]. Remarkably, the CuFe2O4/ZnFe2O4 electrode showed the smallest diameter of the semicircle, indicating a higher charge transfer rate, i.e., smaller charge transfer resistance and, consequently, a smaller recombination rate, suggesting a faster charge transfer and more efficient electron–hole pair separation. This finding suggests that the formation of a barrier potential at the CuFe2O4/ZnFe2O4 interface enhances electron–hole charge transfer.
Consequently, the CuFe2O4/ZnFe2O4CuFe2O4/ZnFe2O4 heterostructure holds promise as a potential photoanode for high-efficiency solar hydrogen generation due to its fast charge separation and transport capabilities.
The Mott–Schottky analysis was conducted to explore the band structure alignment of CuFe2O4 and ZnFe2O4 samples. Their valence and conduction band positions were assessed using the flat-band potential (Vfb) and bandgap values (Eg). Figure 6 depicts the Mott–Schottky (M-S) curves for CuFe2O4 and ZnFe2O4 samples, revealing a negative slope for CuFe2O4 (indicating a p-type semiconductor) and a positive slope for ZnFe2O4 (indicating an n-type semiconductor) [38,39]. The flat-band potential (Vfb) was determined by extrapolating the plot of 1/C2 versus the potential to the x-axis. For ZnFe2O4 and CuFe2O4, the flat-band potentials were −0.74 eV and 0.62 eV vs. Ag/AgCl, respectively. As the difference between the flat-band potential and the bottom/top edge of the conduction/valence band is negligible for n-type/p-type materials, we regard the flat-band potential as the conduction/valence band edge for n-type and p-type semiconductors, respectively [39]. From the flat-band positions and the band energy gaps, the band structure alignment between ZnFe2O4 and CuFe2O4 could be estimated.
Figure 7 depicts the proposed electronic band structure of the p–n junction between CuFe2O4 and ZnFe2O4 under the light condition of our experiment. Strong band bending occurs, leading to the generation of a high barrier potential at the interface. Due to the small thickness of the CuFe2O4 layer on ZnFe2O4, some photons are absorbed within the bulk of ZnFe2O4 (Eg1), while others are absorbed at the surface of CuFe2O4 (Eg2).

3.4. Linear Sweep Voltammetry (LSV) Measurements

A three-electrode configuration was employed using a photoelectrochemical cell (PEC) to investigate the photocurrent density–voltage (J-V) characteristics. The experimental setup is depicted in Figure 8. The modified stainless steel plate served as the working electrode (WE), graphite was used as the counter electrode (CE), and Ag/AgCl served as the reference electrode (RE). J-V measurements were conducted in an alkaline medium (0.3 M NaOH) using a 100-watt white LED source and a concentrated lens under both dark and illuminated conditions. All of the electric components of the experiment (power supply and LED light source) were supplied by a 500-watt solar panel system. The scanning speed rate was fixed at 10 mV/s.
Figure 9 presents the J-V curves for all samples. It is evident that both CuFe2O4 and ZnFe2O4 electrodes exhibited an onset potential of approximately 0.9 volts. Additionally, CuFe2O4 achieved its maximum current at 0.95 volts, while ZnFe2O4 reached its maximum current at 1.01 volts. The decrease in current beyond a certain anode potential is attributed to the onset of electrode oxidation at specific potentials (cathodic potential). Considering that CuFe2O4 possesses two oxidation–reduction mechanisms (Fe2   Fe³⁺ and Cu⁺¹   Cu2⁺), and ZnFe2O4 has only one (Fe2   Fe³⁺), it can be concluded that the corrosion resistance of ZnFe2O4 is superior to that of CuFe2O4.
Moreover, the effect of illumination on the current density was more pronounced for ZnFe2O4 compared to CuFe2O4. This can be explained in terms of the recombination rate between the generated electron–hole pairs under light illumination. As discussed in the section regarding photoluminescence (PL) and electrochemical impedance spectroscopy (EIS) measurements, the PL signal and radius of the semicircles in Nyquist plots for CuFe2O4 were significantly greater than those for ZnFe2O4, indicating a higher recombination rate in CuFe2O4, which in turn affected the photogenerated current. The increase in photocurrent for both Cu ferrite and Zn ferrite was attributed to the enhanced photoconductivity of the samples.
For the p–n junction sample, it is noteworthy that the starting potential was 1.05 volts, and the anodic voltage was 1.57 volts, indicating improved corrosion resistance. Indeed, the performance remains lower than the other samples until reaching 1.44 V (vs. Ag/AgCl) due to the altered bandgap of the CuFe2O4 layer. This alteration causes the band binding for the holes to be greater than that of the electrons (see Figure 7). In other words, the separation rate of generated holes on the Zn ferrite side is faster than the separation rate of generated electrons on the Cu ferrite side, increasing the likelihood of recombination between the holes and electrons. At an applied voltage of 1.44 V (vs. Ag/AgCl), it appears that the rate of electron separation becomes sufficiently fast to overcome recombination with holes. Furthermore, the CuFe2O4 material itself might hinder changes despite the improved charge separation. Additionally, the effect of illumination on the generated current was more pronounced for the p–n junction sample compared to the other electrodes. These results can be explained by considering the suggested p–n junction electronic band diagram, which demonstrates the generation of an internal electric potential. This facilitates the transfer of photogenerated holes to CuFe2O4, while more photogenerated electrons can be transferred to the conductive substrate via the ZnFe2O4 layer. Furthermore, the presence of electron trapping at the interface reduced the recombination rate between the photogenerated electrons and holes, leading to an enhancement in photoelectrochemical performance.
Based on these findings, it is evident that the p–n junction sample, which acts as a photoanode, offers significant advantages for hydrogen generation, particularly under illuminated conditions.

3.5. Evaluation of Hydrogen Production Rate

In this part of the study, the hydrogen production rate (HPR) for the samples was evaluated under both dark and illuminated conditions at a fixed voltage of 1.5 volts (vs. AgCl/Ag) over a period of 10 h. Figure 10 illustrates the relationship between HPR and time. It is apparent that the steady state of production is reached after approximately 1 h of the experiment, and the production rate remains relatively constant throughout the duration of the experiment. This indicates a high corrosion resistance of the modified electrodes. Furthermore, a significant increase in HPR was observed under light illumination.
The HPR under illumination exhibited an enhancement of approximately 30% compared to the measurements taken in the dark. The mechanism of hydrogen production under illumination is depicted in Figure 11.
To examine the effect of photoelectrolysis over 10 h on the crystal and surface structures of the CuFe2O4/ZnFe2O4 p–n junction, XRD and SEM patterns were compared before and after the hydrogen production experiment. Figure 12 and Figure 13 show the XRD chart and the SEM images before and after the experiment, respectively. The XRD chart indicates that the structure remains a single crystal ferrite even after 10 h of experimentation, demonstrating the chemical stability of the sample during prolonged hydrogen production. Furthermore, the SEM images revealed minimal changes in surface morphology; the sample became smoother, except for some etching, particularly at the grain boundaries. This study highlights the good chemical and physical stability of the investigated sample during hydrogen production.
SEM images show a highly rough surface structure, which results from the spray pyrolysis method of preparation, affecting the quality of the prepared samples. For future studies, a new method of thin film preparation, such as atomic layer deposition (ALD), may be used to enhance PEC cell performance.

4. Conclusions

The implementation of a CuFe2O4/ZnFe2O4 heterojunction demonstrated a significant improvement in the hydrogen production rate, achieving an enhancement of approximately 30%. This enhancement can be attributed to the presence of an interfacial barrier potential, which effectively reduces electron–hole recombination rates. Consequently, the photocatalytic effects were more pronounced, leading to improved performance at the electrolyte–electrode interface. Additionally, coating stainless steel with this ferrite compound resulted in enhanced corrosion resistance. The investigated sample exhibits good chemical and physical stability after a relatively long period of hydrogen production.

Funding

This research received no external funding.

Data Availability Statement

Data is contained within the article.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Tolod, K.R.; Hernández, S.; Russo, N. Recent Advances in the bivo4 Photocatalyst for Sun-Driven Water Oxidation: Top-Performing Photoanodes and Scale-Up Challenges. Catalysts 2016, 7, 13. [Google Scholar] [CrossRef]
  2. Ho-Kimura, S.; Moniz, S.J.A.; Handoko, A.D.; Tang, J. Enhanced photoelectrochemical water splitting by nanostructured BiVO4–TiO2 composite electrodes. J. Mater. Chem. A 2014, 2, 3948–3953. [Google Scholar] [CrossRef]
  3. Gholipour, M.R.; Dinh, C.-T.; Béland, F.; Do, T.-O. Nanocomposite heterojunctions as sunlight-driven photocatalysts for hydrogen production from water splitting. Nanoscale 2015, 7, 8187–8208. [Google Scholar] [CrossRef] [PubMed]
  4. Chiarello, G.L.; Selli, E. Photocatalytic hydrogen production. Recent Patents Eng. 2010, 4, 155–169. [Google Scholar] [CrossRef]
  5. Liao, C.-H.; Huang, C.-W.; Wu, J.C.S. Hydrogen Production from Semiconductor-based Photocatalysis via Water Splitting. Catalysts 2012, 2, 490–516. [Google Scholar] [CrossRef]
  6. Moniz, S.J.A.; Shevlin, S.A.; Martin, D.J.; Guo, Z.-X.; Tang, J. Visible-light driven heterojunction photocatalysts for water splitting—A critical review. Energy Environ. Sci. 2015, 8, 731–759. [Google Scholar] [CrossRef]
  7. Zhou, L.; Wang, W.; Liu, S.; Zhang, L.; Xu, H.; Zhu, W. A sono-chemical route to visible-light-driven high-activity BiVO4 photocatalyst. J. Environ. Sci. 2006, 24, 449–457. [Google Scholar] [CrossRef]
  8. Zhang, L.; Li, Y.; Zhang, Q.; Wang, H. Hierarchical nanostructure of WO3 nanorods on TiO2 nanofibers and the enhanced visible light photocatalytic activity for degradation of organic pollutants. CrystEngComm 2013, 15, 5986–5993. [Google Scholar] [CrossRef]
  9. Mallesh, S.; Srinivas, V. A comprehensive study on thermal stability and magnetic properties of MnZn-ferrite nanoparticles. J. Magn. Magn. Mater. 2018, 475, 290–303. [Google Scholar] [CrossRef]
  10. Bastianello, M.; Gross, S.; Elm, M.T. Thermal stability, electrochemical and structural characterization of hydrothermally synthesised cobalt ferrite (CoFe2O4). RSC Adv. 2019, 9, 33282–33289. [Google Scholar] [CrossRef]
  11. Hammad, T.M.; Salem, J.K.; Abu Amsha, A.; Hejazy, N.K. Optical and magnetic characterizations of zinc substituted copper ferrite synthesized by a co-precipitation chemical method. J. Alloy. Compd. 2018, 741, 123–130. [Google Scholar] [CrossRef]
  12. Joshi, G.P.; Saxena, N.S.; Mangal, R.; Mishra, A.; Sharma, T.P. Band gap determination of Ni-Zn ferrites. Bull. Mater. Sci. 2003, 26, 387–389. [Google Scholar] [CrossRef]
  13. Azab, A.A.; Mansour, A.M.; Turky, G.M. Structural, Magnetic, and Dielectric properties of Sr4Fe6O13 ferrite prepared of small crystallites. Sci. Rep. 2020, 10, 4955. [Google Scholar] [CrossRef]
  14. Ismael, M. Ferrites as solar photocatalytic materials and their activities in solar energy conversion and environmental protection: A review. Sol. Energy Mater. Sol. Cells 2020, 219, 110786. [Google Scholar] [CrossRef]
  15. Reddy, D.H.K.; Yun, Y.-S. Spinel ferrite magnetic adsorbents: Alternative future materials for water purification? Coord. Chem. Rev. 2016, 315, 90–111. [Google Scholar] [CrossRef]
  16. ElNahrawy, A.M.; Mansour, A.M.; ElAttar, H.A.; Sakr, E.M.M.; Soliman, A.A.; Hammad, A.B.A. Impact of Mn-substitution on structural, optical, and magnetic properties evolution of sodium–cobalt ferrite for opto-magnetic applications. J. Mater. Sci. Mater. Electron. 2020, 31, 6224–6232. [Google Scholar] [CrossRef]
  17. Jang, Y.J.; Bin Park, Y.; Kim, H.E.; Choi, Y.H.; Choi, S.H.; Lee, J.S. Oxygen-Intercalated CuFeO2 Photocathode Fabricated by Hybrid Microwave Annealing for Efficient Solar Hydrogen Production. Chem. Mater. 2016, 28, 6054–6061. [Google Scholar] [CrossRef]
  18. Jiang, J.; Fan, W.; Zhang, X.; Bai, H.; Liu, Y.; Huang, S.; Mao, B.; Yuan, S.; Liu, C.; Shi, W. Rod-in-tube nanostructure of MgFe2O4: Electrospinning synthesis and photocatalytic activities of tetracycline. New J. Chem. 2016, 40, 538–544. [Google Scholar] [CrossRef]
  19. Guijarro, N.; Bornoz, P.; Prévot, M.; Yu, X.; Zhu, X.; Johnson, M.; Jeanbourquin, X.; Le Formal, F.; Sivula, K. Evaluating spinel ferrites MFe2O4 (M = Cu, Mg, Zn) as photoanodes for solar water oxidation: Prospects and limitations. Sustain. Energy Fuels 2018, 2, 103–117. [Google Scholar] [CrossRef]
  20. Kim, J.H.; Jang, Y.J.; Kim, J.H.; Jang, J.-W.; Choi, S.H.; Lee, J.S. Defective ZnFe2O4 nanorods with oxygen vacancy for photoelectrochemical water splitting. Nanoscale 2015, 7, 19144–19151. [Google Scholar] [CrossRef]
  21. Karim, K.M.R.; Ong, H.R.; Abdullah, H.; Yousuf, A.; Cheng, C.K.; Khan, M.R. Photoelectrochemical reduction of carbon dioxide to methanol on p-type CuFe2O4 under visible light irradiation. Int. J. Hydrogen Energy 2018, 43, 18185–18193. [Google Scholar] [CrossRef]
  22. Tarek, M.; Karim, K.M.R.; Sarkar, S.M.; Deb, A.; Ong, H.R.; Abdullah, H.; Cheng, C.K.; Khan, M.R. Hetero-structure CdS–CuFe2O4 as an efficient visible light active photocatalyst for photoelectrochemical reduction of CO2 to methanol. Int. J. Hydrogen Energy 2019, 44, 26271–26284. [Google Scholar] [CrossRef]
  23. Duan, K.; Que, T.; Koppala, S.; Balan, R.; Lokesh, B.; Pillai, R.; David, S.; Karthikeyan, P.; Ramamoorthy, S.; Lekshmi, I.C.; et al. A facile route to synthesize n-SnO2/p-CuFe2O4 to rapidly degrade toxic methylene blue dye under natural sunlight. RSC Adv. 2022, 12, 16544–16553. [Google Scholar] [CrossRef] [PubMed]
  24. Balatskiy, D.; Budnikova, Y.; Bratskaya, S.; Vasilyeva, M. TiO2-CoFe2O4 and TiO2-CuFe2O4 Composite Films: A New Approach to Synthesis, Characterization, and Optical and Photocatalytic Properties. J. Compos. Sci. 2023, 7, 295. [Google Scholar] [CrossRef]
  25. Liu, Z.; Zhao, Z.-G.; Miyauchi, M. Efficient Visible Light Active CaFe2O4/WO3 Based Composite Photocatalysts: Effect of Interfacial Modification. J. Phys. Chem. C 2009, 113, 17132–17137. [Google Scholar] [CrossRef]
  26. Kim, E.S.; Kang, H.J.; Magesh, G.; Kim, J.Y.; Jang, J.-W.; Lee, J.S. Improved Photoelectrochemical Activity of CaFe2O4/BiVO4 Heterojunction Photoanode by Reduced Surface Recombination in Solar Water Oxidation. ACS Appl. Mater. Interfaces 2014, 6, 17762–17769. [Google Scholar] [CrossRef] [PubMed]
  27. Elzwawy, A.; Mansour, A.; Magar, H.S.; Hammad, A.B.A.; Hassan, R.Y.; El Nahrawy, A.M. Exploring the structural and electrochemical sensing of wide bandgap calcium phosphate/CuxFe3-xO4 core-shell nanoceramics for H2O2 detection. Mater. Today Commun. 2022, 33, 104574. [Google Scholar] [CrossRef]
  28. Ahmed, M.G.; Kandiel, T.A.; Ahmed, A.Y.; Kretschmer, I.; Rashwan, F.; Bahnemann, D. Enhanced Photoelectrochemical Water Oxidation on Nanostructured Hematite Photoanodes via p-CaFe2O4/n-Fe2O3 Heterojunction Formation. J. Phys. Chem. C 2015, 119, 5864–5871. [Google Scholar] [CrossRef]
  29. Lee, S.Y.; Huang, E.-W.; Woo, W.; Yoon, C.; Chae, H.; Yoon, S.-G. Dynamic Strain Evolution around a Crack Tip under Steady- and Overloaded-Fatigue Conditions. Metals 2015, 5, 2109–2118. [Google Scholar] [CrossRef]
  30. Najmoddin, N.; Beitollahi, A.; Kavas, H.; Mohseni, S.M.; Rezaie, H.; Åkerman, J.; Toprak, M.S. XRD cation distribution and magnetic properties of mesoporous Zn-substituted CuFe2O4. Ceram. Int. 2014, 40, 3619–3625. [Google Scholar] [CrossRef]
  31. Tatarchuk, T.; Bououdina, M.; Macyk, W.; Shyichuk, O.; Paliychuk, N.; Yaremiy, I.; Al-Najar, B.; Pacia, M. Structural, Optical, and Magnetic Properties of Zn-Doped CoFe2O4 Nanoparticles. Nanoscale Res. Lett. 2017, 12, 141. [Google Scholar] [CrossRef] [PubMed]
  32. Somnath; Batoo, K.M.; Raslan, E.H.; Adil, S.F.; Sharma, I.; Kumar, G. Correction to: Investigation of electrical, magnetic, and optical properties of silver-substituted magnesium–manganese ferrite nanoparticles. J. Mater. Sci. Mater. Electron. 2020, 31, 10939. [Google Scholar] [CrossRef]
  33. Karnaji; Nurhasanah, I. Photodegradation of Rhodamine B by using ZnFe2O4 Nanoparticles Synthesized through Precipitation Method. IOP Conf. Ser. Mater. Sci. Eng. 2017, 202, 012044. [Google Scholar] [CrossRef]
  34. Park, S.; Baek, J.H.; Zhang, L.; Lee, J.M.; Stone, K.H.; Cho, I.S.; Guo, J.; Jung, H.S.; Zheng, X. Rapid Flame-Annealed CuFe2O4 as Efficient Photocathode for Photoelectrochemical Hydrogen Production. ACS Sustain. Chem. Eng. 2019, 7, 5867–5874. [Google Scholar] [CrossRef]
  35. Kumar, K.V. Tunable Optical Bandgap of Gadolinium Substituted Nickel-Zinc Ferrite Nanoparticles-Effect of Calcination Temperature on Its Optical Parameters. Adv. Mater. Phys. Chem. 2022, 12, 33–45. [Google Scholar] [CrossRef]
  36. Shetty, K.; Renuka, L.; Nagaswarupa, H.P.; Nagabhushana, H.; Anantharaju, K.S.; Rangappa, D.; Prashantha, S.C.; Ashwini, K. A comparative study on CuFe2O4, ZnFe2O4 and NiFe2O4: Morphology, Impedance and Photocatalytic studies. Mater. Today Proc. 2017, 4, 11806–11815. [Google Scholar] [CrossRef]
  37. Belhadj, H.; Messaoudi, Y.; Khelladi, M.R.; Azizi, A. A facile synthesis of metal ferrites (MFe2O4, M = Co, Ni, Zn, Cu) as effective electrocatalysts toward electrochemical hydrogen evolution reaction. Int. J. Hydrogen Energy 2022, 47, 20129–20137. [Google Scholar] [CrossRef]
  38. Preethi, S.; Vivek, S.; Priya, R.; Balakumar, S.; Babu, K.S. Enhanced photocatalytic performance of CuFeO2-ZnO heterostructures for methylene blue degradation under sunlight. J. Mater. Sci. Mater. Electron. 2021, 32, 22256–22269. [Google Scholar] [CrossRef]
  39. Zhang, M.; Xie, X.; Si, Y.; Gao, J.; Du, H.; Pei, S.; Zhang, X.; Yan, Q. Enhanced photocatalytic performance of ZnFe2O4/BiOI hybrid for the degradation of methyl orange. J. Mater. Sci. Mater. Electron. 2019, 30, 8055–8063. [Google Scholar] [CrossRef]
Figure 1. XRD patterns for the prepared samples. * indicates the peaks for stainless steel substrate.
Figure 1. XRD patterns for the prepared samples. * indicates the peaks for stainless steel substrate.
Condensedmatter 09 00031 g001
Figure 2. Diffuse reflection spectra for the prepared samples.
Figure 2. Diffuse reflection spectra for the prepared samples.
Condensedmatter 09 00031 g002
Figure 3. Tauc’s plot analysis indicating the direct bandgap transition(s) for the investigated samples.
Figure 3. Tauc’s plot analysis indicating the direct bandgap transition(s) for the investigated samples.
Condensedmatter 09 00031 g003
Figure 4. Photoluminescence spectra of the prepared samples.
Figure 4. Photoluminescence spectra of the prepared samples.
Condensedmatter 09 00031 g004
Figure 5. Nyquist plot of all investigated samples.
Figure 5. Nyquist plot of all investigated samples.
Condensedmatter 09 00031 g005
Figure 6. Mott–Schottky plots of CuFe2O4 and ZnFe2O4 samples.
Figure 6. Mott–Schottky plots of CuFe2O4 and ZnFe2O4 samples.
Condensedmatter 09 00031 g006
Figure 7. Energy band diagram of CuFe2O4/ZnFe2O4 p–n junction.
Figure 7. Energy band diagram of CuFe2O4/ZnFe2O4 p–n junction.
Condensedmatter 09 00031 g007
Figure 8. Experimental setup for measuring the J-V curve.
Figure 8. Experimental setup for measuring the J-V curve.
Condensedmatter 09 00031 g008
Figure 9. Linear sweep voltammetry (LSV) measurements for all prepared samples.
Figure 9. Linear sweep voltammetry (LSV) measurements for all prepared samples.
Condensedmatter 09 00031 g009
Figure 10. Hydrogen production rate (HPR) at V = 1.5 volts versus time.
Figure 10. Hydrogen production rate (HPR) at V = 1.5 volts versus time.
Condensedmatter 09 00031 g010
Figure 11. Mechanism of hydrogen evolution using CuFe2O4/ZnFe2O4 p–n junction.
Figure 11. Mechanism of hydrogen evolution using CuFe2O4/ZnFe2O4 p–n junction.
Condensedmatter 09 00031 g011
Figure 12. XRD charts for CuFe2O4/ZnFe2O4 junction before and after hydrogen production.
Figure 12. XRD charts for CuFe2O4/ZnFe2O4 junction before and after hydrogen production.
Condensedmatter 09 00031 g012
Figure 13. SEM images of the sample before and after hydrogen production at magnifications of x = 250 and Kx = 1.
Figure 13. SEM images of the sample before and after hydrogen production at magnifications of x = 250 and Kx = 1.
Condensedmatter 09 00031 g013
Table 1. Lattice parameter, crystallite size, and microstrain of the prepared samples.
Table 1. Lattice parameter, crystallite size, and microstrain of the prepared samples.
SampleLattice Parameter (Å)Crystallite Size (nm)Microstrain
ZnFe2O48.425230.005
CuFe2O48.384390.002
CuFe2O4/ZnFe2O48.416470.011
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

Al Turkestani, M.K. Enhancing the Photoelectrochemical Performance of a Superlattice p–n Heterojunction CuFe2O4/ZnFe2O4 Electrode for Hydrogen Production. Condens. Matter 2024, 9, 31. https://doi.org/10.3390/condmat9030031

AMA Style

Al Turkestani MK. Enhancing the Photoelectrochemical Performance of a Superlattice p–n Heterojunction CuFe2O4/ZnFe2O4 Electrode for Hydrogen Production. Condensed Matter. 2024; 9(3):31. https://doi.org/10.3390/condmat9030031

Chicago/Turabian Style

Al Turkestani, M. K. 2024. "Enhancing the Photoelectrochemical Performance of a Superlattice p–n Heterojunction CuFe2O4/ZnFe2O4 Electrode for Hydrogen Production" Condensed Matter 9, no. 3: 31. https://doi.org/10.3390/condmat9030031

Article Metrics

Back to TopTop