Coupling the power of spinel nanoparticles: dual-function metal ferrite catalysts for advanced piezo-photocatalytic and electrocatalytic water splitting

Saima Jabeen a, Mohsin Saleem *a, Farah Mumtaz a, Sofia Javed *a, Mukarram Ali b, Muhammad Zubair Khan *c, Jung-Hyuk Koh *d, Adnan Maqbool e, Abrar H. Baluch f, Muhammad Bilal Khan Niazi g and Iftikhar Hussain h
aSchool of Chemical and Materials Engineering (SCME), National University of Sciences and Technology (NUST), Islamabad 44000, Pakistan. E-mail: mohsin.saleem@scme.nust.edu.pk; mohsin852@cau.ac.kr; sofia.javed@scme.nust.edu.pk
bDepartment of Chemical Engineering, Quantum Nano Center, University of Waterloo, Waterloo, Ontario N2L 3G1, Canada
cDepartment of Materials Science & Engineering, Pak-Austria Fachhochschule, Institute of Applied Sciences and Technology, Mang, Haripur 22621, KPK, Pakistan. E-mail: zubair.khan@fcm3.paf-iast.edu.pk
dSchool of Electrical and Electronics Engineering, Chung-Ang University, Seoul, South Korea. E-mail: jhkoh@cau.ac.kr
eDepartment of Metallurgical and Materials Engineering, University of Engineering and Technology, G.T Road, Lahore, 54890, Pakistan
fDepartment of Aerospace Engineering, & Interdisciplinary Research Center for Aviation and Space Exploration, King Fahd University of Petroleum & Minerals (KFUPM), 31261, Dhahran, Saudi Arabia
gDepartment of Chemical Engineering, & Interdisciplinary Research Center for Refining & Advanced Chemicals, King Fahd University of Petroleum & Minerals (KFPUM), 31261, Dhahran, Saudi Arabia
hDepartment of Mechanical Engineering, City University of Hong Kong, Hong Kong, 999077 China

Received 13th February 2025 , Accepted 6th June 2025

First published on 18th June 2025


Abstract

Hydrogen production through green methods is essential for achieving a green and sustainable energy goal. However, the slow reaction kinetics of the oxygen evolution reaction (OER) and the instability of catalysts for the hydrogen evolution reaction (HER) pose significant challenges to optimizing various water-splitting technologies. Herein, we present a comparative study encompassing mono- and binary metal–organic frameworks (MOFs) and mono- and binary ferrites, where ferrites outperform their MOF counterparts. Both binary MOFs and ferrites exhibited better performance than their mono variants. TEM analysis revealed the nanoscale morphology and crystalline features of the prepared ferrites. At the same time, EDS and BET analysis confirmed the chemical composition, specific surface area, and pore volume, highlighting their structural and electronic properties. Additionally, we compared piezo-photocatalysis and electrocatalysis, providing insights into the smart integration of both technologies. Gas chromatography demonstrated the synergistic coupling of piezoelectricity and photon-excited reactive species in Zn0.5Co0.5Fe2O4. Remarkably, the HER under both sonication and light achieved 484 μmol g−1 h−1, significantly higher than 232 μmol g−1 h−1 and 288 μmol g−1 h−1 for sonication and light alone, respectively. Electrochemical tests confirmed an OER and HER overpotential of 167 and 205 mV, respectively, along with 94% and 73% retention in chronoamperometry, validating its role as a multifunctional water-splitting electrolyzer. This work highlights the potential of designing innovative composites for stable and efficient water-splitting systems such as piezo-photocatalysis and electrocatalysis.


Introduction

Hydrogen has emerged as a pivotal sustainable fuel, with the potential to replace conventional fossil fuels. Its versatility as a key precursor in the synthesis of commodity chemicals, such as green ammonia, and its role in CO2 reduction within a circular economy underscore its significance across a diverse array of sustainable technologies.1–3

The versatility of hydrogen and its high energy density, approximately 120 MJ kg−1, positions it as a crucial sustainable fuel. However, around 90% of commercial hydrogen production currently depends on the steam methane reforming process, pointing out the significant issues of CO2 emissions, facing green credentials and long-term viability concerns. Therefore, there is an urgent need to develop green technologies to address this crucial problem in hydrogen synthesis.4–6

To resolve CO2 emissions and environmental protection, catalysis plays a critical role in the efficient and sustainable advancement of green technology for hydrogen production. H2 generation via the indigenous catalytic water-splitting approach has recently gained a lot of research interest.7 However, the catalytic approach is yet unable to generate H2 on a feasible commercial scale.8 To improve the efficiency and effectiveness of catalytic techniques, recent research is investigating viable catalytic processes, such as electrocatalysis, photocatalysis, and piezo-photocatalysis.9–12 Usually, proton exchange membranes are employed to electrolyze water at room temperature (RT) by using acidic or alkaline solutions. Despite the small-scale efficiency of PEMs being tremendous, the commercial scalability of PEM electrolyzers is still thwarted due to the high cost of catalysts and respective membranes.13–15 Photocatalytic processes, including photocatalysis and using photoelectrochemical cells (PECs), hold significant promise for addressing the energy crisis. However, charge recombination of photogenerated charge carriers significantly hampers photocatalytic efficiency, thereby limiting the commercial potential of these technologies.16–18

Piezo-catalysis harnesses mechanical vibrations to stimulate catalytic reactions and is emerging as a promising technology for efficient hydrogen production.19 Despite its potential to transform the renewable energy landscape, it faces limitations related to the underwater stability of piezo materials and the presence of stationary charge carriers, in contrast to the mobile carriers found in photocatalysis.20,21 Piezo-photocatalysis represents an innovative approach that combines mechanical stress and light irradiation to enhance catalytic activity, offering a promising pathway for efficient energy conversion and environmental remediation.22–24 Unlike standalone photocatalysis, where photogenerated charge carriers recombine before participating in redox reactions, piezo-photocatalysis introduces an internal electric field within the catalyst material due to piezoelectric polarization. This internal field facilitates charge separation by directing electrons and holes in opposite directions, thereby improving charge transport efficiency and reducing recombination losses.25,26 As a result, piezo-photocatalytic systems demonstrate higher quantum efficiencies and superior hydrogen production rates compared to traditional photocatalysts.

Strain-induced piezoelectricity in nanomaterials often results in poor electrical conductivity and a reduced concentration of free charge carriers, which hinders their ability to drive electrochemical reactions.17,27 Similarly, high charge recombination in photocatalysis limits the straightforward application of these materials in water splitting. However, combining piezoelectricity with photocatalytic nanomaterials can effectively address the inherent limitations of both processes. Forced polarization from piezo materials decreases charge recombination in photocatalysts, thereby enhancing overall catalytic efficiency.28,29 Additionally, the higher conductivity of photocatalysts, when paired with piezo materials, increases the overall charge transport on the surface and bulk of photocatalysts, thus boosting the HER and OER rates.30–34 The internal piezoelectric field synergistically interacts with photogenerated carriers, accelerating their transfer to reactive sites. This effect has been observed to boost charge separation efficiency by 1.5x to 3x compared to conventional photocatalysis. Additionally, this mechanism enables better utilization of photogenerated carriers, reducing energy losses and improving the overall catalytic performance.35 Once the charge carriers reach the catalytic surface, they participate in redox reactions. Electrons reduce protons (H+) to generate H2, while holes oxidize water molecules to produce O2. The piezoelectric effect also contributes to modulating the band structure, further enhancing photon absorption and improving charge transport dynamics at the catalyst surface.36

Recent studies have demonstrated that applying sound waves at specific frequencies (20–40 kHz) introduces additional mechanical stress in piezoelectric materials, significantly enhancing charge carrier separation efficiency. This effect is attributed to the increased piezoelectric polarization, which acts synergistically with light irradiation to suppress charge recombination by up to 40%.37,38 In ZnO-based piezo-photocatalysts, ultrasound-assisted hydrogen evolution rates have been reported to improve by a factor of 2.3x compared to conventional photocatalysis. Similarly, BaTiO3-based composites have achieved an HER rate of 6.1 mmol g−1 h−1 under simultaneous piezoelectric and photocatalytic activation.39,40 These findings highlight the potential of integrating acoustic stimulation into piezo-photocatalytic systems to achieve higher efficiency in water splitting.

A variety of materials have been used in piezo-photocatalysis ranging from classical metal oxide, nitride, and sulfide semiconductors and titanates. Their piezoelectric features emerge from their non-centrosymmetric nature including ZnO, GaN, InN, CdS, MoS2, BaTiO3, BiTiO3, BiNaTiO3, PbZrTiO3, and any composite of these materials.41–44 Similar types of oxides, sulfides, and nitrides have also been used in electrocatalysis and photo-electrocatalysis in the quest of finding an economical yet efficient catalyst.45,46 Materials with borderline piezoelectric properties like ferrites such as BiFeO3 and titanates such as BiNaTiO3 are ideal for water-splitting applications via both piezo-photocatalytic and electrocatalytic routes.20,47,48 Similarly, metal–organic frameworks (MOF) and covalent organic frameworks (COF) nanomaterials are ideal not only for their hydrogen evolution reaction (HER) but also for their versatility to be used in both processes like borderline piezoelectric materials.49–51

Metal–organic frameworks (MOFs) have gained attention as advanced materials in piezo-photocatalysis due to their extended polymeric chains, which enhance the stability and longevity of metal nanoparticles by preventing underwater oxidation.52 Similarly, perovskite halides have seen recent advancements through the incorporation of non-metal co-catalysts, although research on co-catalysts and promoters in piezo-photocatalysis is still in its infancy.53 At the same time, hybrid techniques are revolutionizing water-splitting technologies. The integration of triboelectric and piezoelectric nanogenerators with photovoltaic systems in photoelectrochemical cells eliminates the need for external bias, a significant limitation of conventional methods.54,55 These innovations not only improve efficiency but also create opportunities for coherence among methods such as photocatalysis, piezocatalysis, piezo-photocatalysis, and photoelectrochemical processes. The combination of novel materials and hybrid techniques underscores the growing potential to address challenges in renewable energy and catalysis.56,57

Ferrites have been extensively studied in electro-catalysis, but their potential in piezo-photocatalysis remains largely unexplored. Among ferrites, BiFe2O3 is the only material investigated for piezo-photocatalytic applications. Given the borderline properties of ferrites that are akin to those of metals and piezoelectrics, a broader range of ferrite materials may be well-suited for piezo-photocatalytic water splitting.52,56,57 In electro-catalysis, the metal component of ferrites serves as the active site for the hydrogen evolution reaction (HER) while oxygen vacancies act as active sites for the oxygen evolution reaction (OER). Similarly, in metal–organic frameworks (MOFs), metal ions contribute to HER sites and organic linkers serve as active sites for the OER. Due to the bifunctional potential of both MOFs and ferrites, we have explored their water-splitting capabilities in two distinct processes.58,59

We have conducted an in-depth study comparing two distinct material types, i.e. metal ferrites and MOFs, to investigate water splitting in both piezo-photocatalysis and electrocatalysis. Our main goal in this research is to develop the zinc–cobalt ferrite for water splitting by using piezo-photocatalysis and electrocatalysis. According to our findings in both mono and binary variants of the MOFs and ferrites, the binary variants perform better in water splitting by using piezo-photocatalysis materials in both processes than their mono counterparts.

Experimental methods

Materials

All chemicals were purchased from Sigma-Aldrich and used without additional purification such as iron nitrate nonahydrate (Fe(NO3)3·9H2O), zinc nitrate tetrahydrate (Zn(NO3)2·4H2O), cobalt nitrate hexahydrate (Co(NO3)2·6H2O, sodium hydroxide (NaOH) and ethanol. Additionally, a Philips 635 nm bulb is employed in the reaction setup.

Synthesis of piezo-photocatalysts

As shown in Fig. 1, the synthesis of Zn0.5Co0.5Fe2O4 nanoparticles is carried out using the co-precipitation method, a widely recognized technique for producing nanoparticles with controlled properties. The process begins by dissolving 16.00 grams of iron(III) nitrate nonahydrate, 1.45 grams of cobalt(II) nitrate hexahydrate, and 1.45 grams of zinc(II) nitrate tetrahydrate in 400 milliliters of deionized water to prepare a precursor solution. The reaction is conducted in a glass beaker ensuring chemical stability and preventing contamination. Sodium hydroxide solution is added dropwise to the solution while stirring vigorously at a speed maintained between 1500 and 2500 rpm until the pH reaches 13. The reaction is maintained at 80 °C for an hour. After the reaction, the nanoparticles are cooled, washed with deionized water and ethanol through centrifugation, and then sintered at 800 °C for three hours to consolidate the material and improve its properties.
image file: d5ta01180a-f1.tif
Fig. 1 Schematic of the co-precipitation synthesis of ZnCoFe2O4 nanoparticles.

A similar co-precipitation method is employed for synthesizing CoFe2O4 and ZnFe2O4, where 2.91 grams of cobalt(II) nitrate hexahydrate and zinc(II) nitrate tetrahydrate are incorporated into the precursor solution respectively. The remainder of the process, including purification and sintering, mirrors the procedure used for Zn0.5Co0.5Fe2O4 in Fig. S1. All three synthesized materials, CoFe2O4, ZnFe2O4, and Zn0.5Co0.5Fe2O4, are essential for further characterization and potential applications in water splitting and hydrogen production.

Catalyst performance measurements

A Sigma-Aldrich Quickfit 24/29 flask with a rubber septum was used for the gas collection system, and a 10 mL gas-tight syringe was used to extract the H2 sample to prevent leakage. To ensure accuracy, the system was pre-tested with N2 gas to confirm no loss before actual measurements. Photo illumination was provided using a Philips Master HPI-T Plus 400W/645 E40 neutral stable white multi-gear bulb positioned at a fixed distance from the reaction flask to maintain consistent light intensity.60 The sonicator, along with the reaction flask, was placed inside the photo chamber maintained at 25 °C. A Shimadzu GC-2010 Pro gas chromatograph, equipped with a thermal conductivity detector, measured the gas percentages three times from 10 mL samples of the reaction mixtures to improve precision. H2 moles were calculated using the standard molar volume formula n = V/Vm.

Characterization

The characterization of the synthesized ferrite materials involves several key techniques. X-Ray Diffraction (XRD) is used for phase identification and crystallite size calculation by analyzing the diffraction patterns of finely ground ferrite powders. Structural parameters were analyzed using Rietveld refinement of the powder diffraction patterns with the help of CASAS software, and the refined crystal structures were visualized using VESTA software. Field Emission Transmission Electron Microscopy (FETEM) and Scanning Electron Microscopy (SEM) provide detailed images of the surface morphology and particle size after the ferrite powder is dispersed in ethanol, ultrasonicated, and drop-cast onto a conductive substrate. Fourier-Transform Infrared Spectroscopy (FTIR) is used to identify functional groups and bonding characteristics.

Electrochemical analysis

For electrochemical analysis, electrodes were fabricated by mixing 2 mg of the synthesized electrocatalyst with Nafion as a binder and carbon black to enhance conductivity. The resulting slurry was uniformly applied onto nickel foam (NF), which served as the working electrode. The electrocatalytic performance of CoFe2O4, ZnFe2O4, and ZnCoFe2O4 electrodes was evaluated for both the HER and OER in a 6 M potassium hydroxide (KOH) aqueous electrolyte. The electrochemical measurements were conducted at room temperature using a standard three-electrode cell configuration. A platinum wire was used as the counter electrode, while an Ag/AgCl electrode served as the reference. All potentials were converted to reversible hydrogen electrode (RHE) values, with the formula E(RHE) = E (Ag/AgCl) + 0.059 × pH. The overpotentials were appropriately adjusted for IR compensation to account for ohmic losses.61,62

Results and discussion

Fig. 2(a, b, and c), S2 and Table 1 provide a comprehensive structural analysis of the spinel ferrite nanoparticles ZnFe2O4, CoFe2O4, and CoZnFe2O4. The X-ray diffraction (XRD) patterns confirm the cubic spinel structure having a space group of Fd3m,63 with lattice parameters derived as a = 8.429 Å for ZnFe2O4, a = 8.371 Å for CoFe2O4, and a = 8.412 Å for CoZnFe2O4 as shown in Fig. 3(b). These values align with the observed shifts in peak positions, reflecting the substitution effects of Co2+ and Zn2+ ions due to their ionic radii (Co2+: 0.72 Å and Zn2+: 0.74 Å). The slight lattice contraction in CoFe2O4 compared to ZnFe2O4 is consistent with the smaller ionic radius of Co2+, while CoZnFe2O4 exhibits intermediate behavior due to partial substitution. The refinement parameters (χ2) show improved fitting quality with CoZnFe2O4 (2.755) compared to ZnFe2O4 (4.412) and CoFe2O4 (3.517), indicating reliable crystallographic Rietveld modeling. The goodness-of-fit parameter (RWP) follows a similar trend, with CoZnFe2O4 achieving the lowest value (2.114), further confirming the accuracy of its structural refinement. The full width at half-maximum (FWHM) values increase for CoZnFe2O4 (0.561) compared to ZnFe2O4 (0.481) and CoFe2O4 (0.407), indicating smaller crystallite size and higher micro-strain. The Rietveld refinement data were effectively correlated with crystallite size and micro-strain, as shown in Fig. 2 and Table 1. These structural parameters influence defect density, surface area, and charge separation efficiency. Such factors are crucial for enhancing the photocatalytic activity of ferrites in hydrogen production. Therefore, crystallographic analysis directly supports the material's performance validation.
image file: d5ta01180a-f2.tif
Fig. 2 (a) X-ray diffraction (XRD) patterns and schematic crystal structures of ZnFe2O4, CoFe2O4, and ZnCoFe2O4 nanoparticles. Rietveld analysis of the samples demonstrates precise peak fitting and quantitative phase composition, which confirms high crystallinity and structural refinement. (b) The extrapolation of the calculated lattice parameter from each diffraction peak versus the Nelson-Riley function estimates the precise lattice constant of the sample, for which the function is zero. The absence of impurity peaks indicates high phase purity, while slight shifts in peak positions in ZnCoFe2O4 reflect lattice distortion due to Co2+ and Zn2+ substitution. (c) The schematic representations illustrate the spinel crystal lattice with tetrahedral (A-site) and octahedral (B-site) cation arrangements, showcasing structural changes induced by ionic substitution.
Table 1 Summary of lattice parameters of ZnFe2O4, CoFe2O4, and ZnCoFe2O4 nanoparticles
ZnFe2O4 CoFe2O4 CoZnFe2O4
Lattice parameter, a (Å) 8.429 8.371 8.412
χ 2 4.412 3.517 2.755
R WP 3.204 2.259 2.114
FWHM 0.481 0.407 0.561
Crystallite size (nm) 3.023 4.178 3.532
Dislocation density (nm−2) 0.109 0.057 0.080



image file: d5ta01180a-f3.tif
Fig. 3 Morphological and elemental analyses of ZnFe2O4, CoFe2O4, and ZnCoFe2O4 nanoparticles. SEM micrographs in (a), (b), and (c) represent the surface morphology of ZnFe2O4, CoFe2O4, and ZnCoFe2O4, respectively, showing agglomerated spherical nanoparticles.

The crystallite sizes are calculated to be 3.023 nm for ZnFe2O4, 4.178 nm for CoFe2O4, and 3.532 nm for CoZnFe2O4.64 The dislocation density, inversely proportional to crystallite size, is highest for ZnFe2O4 (0.109 nm−2), followed by CoZnFe2O4 (0.080 nm−2) and CoFe2O4 (0.057 nm−2), indicating that CoFe2O4 possesses the most crystalline structure with fewer defects.65

These structural results demonstrate the impact of ionic substitution on lattice parameters, crystallite size, and dislocation density, which directly influence the structural and functional properties of the spinel ferrite nanoparticles in piezo-photocatalysis and electrochemical properties. Also, the co-precipitation method has effectively controlled the synthesis parameters with uniform crystallite size, ensuring high phase purity and tunable crystallographic properties, essential for their potential applications in magnetic, catalytic, and energy storage technologies.

The morphology and composition of ZnCoFe2O4, CoFe2O4, and ZnFe2O4 nanoparticles, as illustrated in Fig. 3, were investigated using FE-SEM and EDS analyses. The SEM micrographs (a), (b), and (c) reveal the presence of spherical nanoparticles with an average size in the range of 20–30 nm. These nanoparticles exhibit slight agglomeration due to their high surface energy and dipole–dipole interactions among the particles, a characteristic commonly observed in nanoscale materials.66

Fig. 3 shows the distinct crystalline grains despite porosity and agglomeration, and the distribution of grain size difference is observable. The existence of porosity and agglomeration could result from the synthesis process, resulting in material characteristic degradation, such as density and surface area.67 The size and morphology of the grains were affected by the substitution of Zn2+ and Co2+ ions. The replacement/substitution of cations is known to change the nucleation and grain growth kinetics, resulting in the structural integrity and behavior of the particles. The SEM analysis highlights that all three ferrite samples share a similar morphology, consistent with their spinel structure. The ZnCoFe2O4 sample shows slightly more pronounced aggregation, which may be attributed to the mixed-ion substitution affecting the interparticle interactions and surface energy balance. This suggests that cation substitution not only alters the crystal structure but also significantly impacts particle packing and agglomeration.67 Corresponding histograms have been included to quantify the particle size distribution for ZnFe2O4, CoFe2O4, and ZnCoFe2O4, providing a clear statistical representation and distribution profile (ESI Fig. S4). The histograms clearly illustrate the particle size uniformity and distribution range, thereby enhancing the clarity and reliability of morphological analyses.

The EDS spectra shown in Fig. S3(e), (f), and (g) further confirm the elemental composition of the synthesized nanoparticles. The presence of oxygen (O), iron (Fe), cobalt (Co), and zinc (Zn) peaks validates the successful incorporation of these elements into the spinel lattice.68 The quantitative analysis of the EDS data provides the weight and atomic percentages of the elements, confirming the stoichiometric composition of ZnCoFe2O4, CoFe2O4, and ZnFe2O4. For ZnCoFe2O4, the simultaneous presence of both Zn and Co signals the effective doping of cobalt into the zinc ferrite structure, which is consistent with the observed lattice distortion and peak shifts in the XRD data in Fig. 2. These analyses confirm the high quality and purity of the sample and the effective integration of transition metal ions into the spinel lattice, resulting in a strong foundation for piezo-photocatalysis in water splitting and electrocatalysis processes.69

Fig. 4 shows the TEM images of the synthesized samples and the discrete morphological characteristics of all the samples. The TEM images provide a deeper insight into particle morphology and uniform dispersion. ZnFe2O4 exhibits a uniform distribution of the particles, while others show large and more agglomerated particles due to the magnetic effect. The particles are well dispersed and uniformly distributed with mainly spherical and irregular shapes and range in the average particle size from 20 to 25 nm. Fig. 4(b) displays slightly larger particles compared to ZnFe2O4, with a mix of spherical and semi-cubic morphologies. Some degrees of particle agglomeration are observed in CoFe2O4 due to magnetic interactions. As shown in Fig. 4(a), ZnFe2O4 exhibits a more uniform particle size distribution with cubic and truncated morphologies, indicating that the integration of Zn and Co into the spinel structure improves particle uniformity.


image file: d5ta01180a-f4.tif
Fig. 4 FETEM images of (a) ZnCoFe2O4, (b) CoFe2O4, and (c) ZnFe2O4 nanoparticles. The TEM micrographs illustrate uniform particle distributions and nanoscale grain sizes ranging from 20 to 25 nm. FETEM images reveal well-defined lattice fringes with inter-planar spacings of 0.26 nm and 0.25 nm corresponding to the (311) planes, 0.21 nm corresponding to the (400) plane, and 0.30 nm corresponding to the (220) plane, confirming the spinel crystal structure of the synthesized samples.

FETEM images further confirm the high crystallinity and well-defined lattice structures of all three samples. In ZnFe2O4, lattice fringes with spacings of 0.26 nm and 0.21 nm correspond to the (311) and (400) planes, respectively, while an additional spacing of 0.30 nm is attributed to the (220) plane. CoFe2O4 displays lattice fringes with a prominent spacing of 0.25 nm for the (311) planes, along with similar spacings of 0.21 nm and 0.30 nm for the (400) and (220) planes, consistent with its spinel structure. The ZnCoFe2O4 sample exhibits lattice fringes with spacings of 0.26 nm and 0.25 nm for the (311) planes and 0.21 nm and 0.30 nm for the (400) and (220) planes.70,71 These observations confirm the successful incorporation of Zn and Co into the spinel matrix without disrupting the structural integrity of the lattice.72 The TEM and FETEM analyses collectively demonstrate the nanoscale size, uniformity, and crystallinity of the synthesized samples, highlighting their suitability for advanced applications.73 These inter-planar spacings and their associated lattice planes in ZnCoFe2O3 align well with the known structural parameters of pristine ZnFe2O4 and CoFe2O4 nanospinels.74

As shown in Fig. 5, TEM-EDS analysis and elemental mapping for ZnCoFe2O4 nanoparticles provide a comprehensive understanding of the material's composition and elemental distribution. The grayscale TEM image shows the nanoparticle morphology, while the overlaid elemental mapping displays the spatial distribution of Zn, Co, Fe, and O. The uniform and overlapping intensities of these maps confirm the homogeneous distribution of all elements, indicating successful integration of Zn and Co into the Fe-based spinel structure. The EDS spectrum highlights characteristic peaks for O, Fe, Co, and Zn, affirming the elemental presence in the sample. Quantitative analysis reveals that Fe accounts for the highest atomic percentage (52.05%), followed by Zn (24.53%) and Co (20.84%), with O making up 2.57%. This composition aligns closely with the theoretical stoichiometry of ZnCoFe2O4, indicating high-purity synthesis and structural consistency. The low oxygen content further suggests minimal surface oxidation, supporting the formation of a stable spinel phase. The TEM-EDS and elemental mapping data for CoFe2O4 and ZnFe2O4 are provided in the ESI file for reference in Fig. 5S(b and c).


image file: d5ta01180a-f5.tif
Fig. 5 (a): FETEM-EDS and elemental mapping data for ZnCoFe2O4 nanoparticles. The FETEM-EDS elemental mapping confirms the uniform spatial distribution of Zn, Co, Fe, and O across the nanoparticle cluster.

Additionally, UV-vis analysis was performed and Tauc plots were obtained to determine the band gap of the ZnCoFe2O4 composite, which was found to be 2.4 eV in Fig. 6(a), which aligns with the existing literature.75,76Fig. 6(b) shows the obtained FTIR spectra of the ZnCoFe2O4, CoFe2O4, and ZnFe2O4 nanoparticle ferrite samples, revealing three primary peaks that provide insights into the vibrational behavior of chemical bonds within the ferrites. First, a notable peak positioned at approximately 3417 cm−1 present in all samples signifies the stretching vibration of O–H bonds present in water molecules.77,78 This common peak is observed consistently across all the samples, attributed to their exposure to ambient air before measurement. The peak at 1618–1632 cm−1 corresponds to the stretching vibration of the C–O bond. Additionally, peaks appearing at 1101–1113 cm−1 are assigned to the water molecules retained in the prepared nanoparticles.79,80 The characteristic peak centered at around 572 cm−1 corresponds to the stretching vibration of M–O bonds at the tetrahedral sites, a key feature of spinel ferrites. This peak's frequency subtly differs among the samples, being slightly higher in ZnCoFe2O4 compared to CoFe2O4 and ZnFe2O4. This disparity implies enhanced Fe–O bond strength in the former samples due to the greater electronegativity of cobalt and ions, resulting in a stronger attraction of oxygen atoms.81


image file: d5ta01180a-f6.tif
Fig. 6 (a) UV-vis spectra and the Tauc plot of ZnCoFe2O4, (b) FTIR spectra and (c) BET surface area of ZnFe2O4, CoFe2O4, and ZnCoFe2O4 nanoparticles showing characteristic O–H stretching (∼3417 cm−1), metal-oxygen vibrations (∼579 cm−1), and other functional group absorptions. The spectra confirm the spinel ferrite structure and reflect cation substitution effects. (c) BET surface area of ZnCoFe2O4.

The similarity in spectra across all three samples confirms their spinel ferrite structure, while the slight differences in peak positions highlight the impact of cation substitution. The main difference is in the tiny spaces within the structure where metal ions sit. In ZnFe2O4, zinc ions exclusively occupy tetrahedral and octahedral sites, while in ZnCoFe2O4, a mix of zinc and cobalt ions influences the Fe–O bond dynamics. These changes in cation composition and their effect on vibrational properties demonstrate how cation substitution affects the local bonding environment and structural behavior of the nanoparticles.81

This difference in the metal ions in these spaces affects the strong bonds between iron (Fe) and oxygen (O) atoms. This, in turn, affects the shifting of the peaks we see in the FTIR measurements. That's why we see slightly higher peaks in ZnCoFe2O4 spectra compared to CoFe2O4 and ZnFe2O4. It's like a small bump in the harmony of the spectra.82 This FTIR analysis confirms the synthesis and structural integrity of the ferrites, providing valuable insights into their chemical and vibrational characteristics.

The BET surface area analysis was conducted for ZnCoFe2O4 using a Micromeritics Gemini V11 2390 model as shown in Fig. 6(c). The samples were preheated at 200 °C for 12 hours to ensure proper degassing. The incorporation of Zn and Co into Fe2O4 resulted in a significant increase in surface area of 151.5 m2 g−1, making the surface nearly twice as active for catalytic applications. The increased surface area enhances accessibility to active sites, thereby improving catalytic efficiency, as catalytic reactions predominantly occur at the material's surface. This increase in surface area aligns with gas chromatography data, which showed a higher hydrogen yield for ZnCoFe2O4, compared to the individual CoFe2O4 and ZnFe2O4 samples.83 The addition of Zn and Co in Fe2O3 increases the surface area by 151.5 m2 g−1i.e. making the overall surface more active for improved catalysis. The catalysis predominantly takes place at the surface, and an increased surface area facilitates greater accessibility to active sites for catalytic reactions.

Additionally, the BJH (Barrett–Joyner–Halenda) adsorption analysis summarized in Fig. 6(g) highlights the accumulative pore area, pore volume, and average pore diameter of ZnCoFe2O4. These parameters further confirm that the incorporation of Zn and Co enhances the textural properties of ZnCoFe2O4 making it more suitable for catalytic reactions by increasing the availability and accessibility of active sites.

To evaluate the piezo-photocatalytic effects on hydrogen production, the H2 evolution performance of the synthesized samples was investigated under both light irradiation and sonication. A 100 mL reaction mixture was prepared with distilled water (DI), incorporating 25 mg of the catalyst.

Fig. 7 provides a comprehensive analysis of the hydrogen (H2) production performance of ZnCoFe2O4, CoFe2O4, and ZnFe2O4 catalysts under various conditions, highlighting their efficiency, stability, and comparative advantages. Fig. 7(a) and (b) demonstrate the H2 production rates of the three samples, with ZnCoFe2O4 achieving the highest rate of 434 μmol g−1 h−1, compared to 412 μmol g−1 h−1 for CoFe2O4 and 268 μmol g−1 h−1 for ZnFe2O4. The accumulated H2 yield further emphasizes the superior catalytic efficiency of Zn0.5Co0.5Fe2O4, attributed to the synergistic role of cobalt and zinc ions in the spinel structure. Fig. 7(c) investigates the individual and combined effects of piezoelectric and photocatalytic mechanisms on ZnCoFe2O4. Sonication alone (piezoelectric effect) produced 232 μmol g−1 h−1 of H2, while light irradiation (photocatalysis) resulted in a slightly higher yield of 288 μmol g−1 h−1. However, when sonication and light were applied simultaneously, synergistic interaction significantly enhanced the H2 production rate to 494 μmol g−1 h−1. Fig. 7(d) examines the cycling stability of ZnCoFe2O4 over seven cycles of one-hour durations. The catalyst maintained stable performance with a slight increase in H2 production from 418 μmol g−1 h−1 in the first cycle to 460 μmol g−1 h−1 in the seventh, demonstrating excellent durability for prolonged use. Fig. 7(e) compares the H2 production efficiency of ZnCoFe2O4 with that of other materials, including Cu-MOF, Fe-MOF, and Mn-MOF. A ZnCoFe2O4 outperformed these materials, showcasing its superior catalytic activity and stability in hydrogen generation. Finally, Fig. 7(f) illustrates the cumulative hydrogen yield of Zn0.5Co0.5Fe2O4 under combined piezoelectric and photocatalytic conditions. The consistent increase in H2 yield over time confirms the catalyst's high efficiency and sustainability for hydrogen evolution. Collectively, these results underline the exceptional performance of ZnCoFe2O4, highlighting its potential as a durable and efficient material for sustainable hydrogen production technologies. A summary of the hydrogen production rates for all tested samples is provided in Table 2. Fig. 7(f) illustrates the cumulative hydrogen yield of ZnCoFe2O4 under simultaneous piezoelectric and photocatalytic conditions, showing a steady increase in H2 production over time. This enhanced yield is attributed to the synergistic interaction between Zn2+ and Co2+ ions within the spinel lattice, which optimizes charge separation and reduces recombination, as supported by recent studies. The dual activation via light and sonication promotes band edge modulation and polarization-induced charge generation, significantly improving catalytic activity.83 This sustained performance also reflects the structural stability of ZnCoFe2O4 over extended cycles, as demonstrated in Fig. 7(d), aligning with literature reports on spinel ferrites exhibiting superior durability in hybrid catalysis systems. ZnCoFe2O4 is not only conductive and photo-piezoactive but also fully capable of generating charges under induced polarization and light irridiation. ZnCoFe2O4 directly splits water into its constituent gases i.e. H2 and O2 by following a series of reactions:

ZnCoFe2O4 ⇌ ZnCoFe2O4 + h+ + e

H2O + h+ ⇌ 2H+ + 1/2O2

2H+ + 2e ⇌ H2


image file: d5ta01180a-f7.tif
Fig. 7 (a) Best compositions for H2 yield among ZnFe2O4, CoFe2O4, and ZnCoFe2O4, highlighting ZnCoFe2O4 as the most efficient. (b) Accumulated H2 amount for each composition under identical conditions. (c) Contributions of photo, piezo, and their synergistic effects to H2 evolution, demonstrating the combined enhancement. (d) Cycling stability of the catalysts, indicating the durability of ZnCoFe2O4 over multiple cycles. (e) Comparative analysis of H2 yield for ZnCoFe2O4versus other reported materials, showcasing its superior performance.
Table 2 The hydrogen production of ZnFe2O4, CoFe2O4, and ZnCoFe2O4 samples
Catalysts & solvents H2 μmol g−1 h−1
Cu-MOF 103
Fe-MOF 107
Mn-MOF 282.7
Cu/Mn-MOF 280.86
Fe/Mn-MOF 308
ZnFe2O4 288
CoFe2O4 412.7
ZnCoFe2O4 434.5
Piezo-catalysis 232
Photo-catalysis 288
Piezo-photocatalysis 484


Reactions were performed using the catalytic water mixture in the dark without sonication (Fig. 8(a)), in the dark with sonication (Fig. (b)), under light without sonication (Fig. (c)), and with both light and sonication (Fig. (d)) and samples collected were subject to GC and the results are shown in Fig. 7(c). The reaction in Fig. 8(a) results in zero HER; however, their synergistic effect (Fig. 8(d)) is evident i.e. 484 μmol g−1 h−1 much more than that of individual catalytic processes i.e. 232 μmol g−1 h−1 for piezocatalysis and 288 μmol g−1 h−1 for photocatalysis alone. Moreover, ZnCoFe2O4 turned out to be more photoactive as compared to its piezoactivity. Herein, forced polarization through piezoelectricity is responsible for reduced charge recombination and aiding photocatalysis for synergistically enhanced HER activity as illustrated in Fig. 8.


image file: d5ta01180a-f8.tif
Fig. 8 Schematics of piezo- photocatalysis, (a) reaction performed in the dark without sonication results in no water splitting reaction demonstrated by zero HER, (b) & (c) piezo- and photocatalytic water splitting respectively, and (d) highest HER activity as a result of the synergistic effect between sonic and photonic energies.

The electrochemical performance of the synthesized ZnCoFe2O4, CoFe2O4, and ZnFe2O4 electrocatalysts, deposited onto nickel foam, was evaluated using linear sweep voltammetry (LSV). The HER and OER were analyzed under identical conditions, employing a scan rate of 20 mV s−1. The potential range was set from 0 to 1.4 V and 0 to −1.4 V for OER and HER measurements respectively. The kinetic performance of electrocatalysts was evaluated by calculating the Tafel slope using the equation (η = a + b  log  j), where η represents the overpotential and b is the Tafel slope. The study employed electrochemical impedance spectroscopy (EIS) to measure long-term stability at a constant applied voltage of 0.6 V. EIS was carried out in the frequency range of 20 kHz to 0.1 Hz. The Ag/AgCl reference electrode potentials were converted to reversible hydrogen electrode (RHE) potentials.84

Fig. 9(a and b) presents comparative OER polarization curves for ZnCoFe2O4, CoFe2O4 and ZnFe2O4 electrocatalysts. The Tafel slope analysis in Fig. 9(a) further underscores the superior catalytic performance of ZnCoFe2O4, with a Tafel slope of 167 mV dec−1, indicating faster electron transfer kinetics. By comparison, CoFe2O4 and ZnFe2O4 exhibit higher Tafel slopes of 260 mV dec−1 and 300 mV dec−1, respectively, which are indicative of slower reaction kinetics. These results suggest that the substitution of Zn and Co into the ferrite lattice improves the catalytic efficiency of ZnCoFe2O4 by lowering the energy barrier for the OER process. The fluctuations in current density observed in the region from approximately 1.4 to 1.6 V vs. RHE are attributed to transient phenomena, specifically bubble nucleation and release due to rapid oxygen evolution at the electrode–electrolyte interface. At elevated potentials, rapid gas evolution (O2 during the OER) can lead to the accumulation and coalescence of gas bubbles on the electrode surface. These bubbles can temporarily block active sites, causing transient increases and decreases in current density due to fluctuating effective surface area and mass transport limitations.85–88 As shown in Fig. 9(b), ZnFe2O4 exhibited higher overpotential and lower catalytic activity compared to ZnCoFe2O4 and CoFe2O4, both of which demonstrated a reduction in activation energy. In the polarization curves, CoFe2O4 and ZnFe2O4 exhibit higher overpotential and the lowest catalytic activity, highlighting their limited effectiveness as OER catalysts. Conversely, ZnCoFe2O4 and CoFe2O4 show significantly reduced overpotential, indicating a notable reduction in activation energy. Among these, ZnCoFe2O4 displayed enhanced OER activity with reduced overpotential outperforming the pristine ferrites, reaching a current density of 10 mA cm−2 at an overpotential of 415 mV. Additionally, the nickel foam substrate exhibited negligible activity for the OER, indicating that its contribution to the overall water-splitting process is minimal. The corresponding Tafel slopes, calculated from the polarization data and presented in Fig. 9(c), are 205 mV dec−1 for ZnCoFe2O4, 228 mV dec−1 for CoFe2O4, and 307 mV dec−1 for ZnFe2O4. A lower Tafel slope indicates more efficient electron transfer kinetics during the HER process, further confirming the superior catalytic properties of ZnCoFe2O4. The reduced Tafel slope for ZnCoFe2O4 suggests a facilitated reaction pathway and a lower energy barrier for hydrogen production. Similarly, the hydrogen evolution reaction (HER) polarization curves, as shown in Fig. 9(d), compare the catalytic performance of ZnCoFe2O4, CoFe2O4, and ZnFe2O4. Among the samples, ZnCoFe2O4 exhibits superior catalytic activity at lower overpotentials compared to the other ferrites. The improved performance of ZnCoFe2O4 is attributed to the synergistic interaction of zinc and cobalt ions within the spinel lattice. Generally, the reaction pathway for the HER in alkaline media undergoes a three-step process, and metal ions in ferrites act as an active site for the HER. Therefore, optimizing metal sites has a direct impact on HER efficiency.89 Water splitting undergoes a series of reactions.


image file: d5ta01180a-f9.tif
Fig. 9 (a) Polarization curves of ZnFe2O4, CoFe2O4, and ZnCoFe2O4, showing the lowest overpotential for ZnCoFe2O4. (b) Tafel slopes highlight its superior kinetics. (c) Polarization curves under varying conditions confirm ZnCoFe2O4 higher efficiency (205 mV). (d) Tafel slopes further demonstrate enhanced kinetic performance. (e and f) Chronoamperometric curve of ZnCoFe2O4 for 14 hours at a constant applied voltage of 0.6 V, showing 94% and 73% retention of current density, indicating excellent long-term stability and durability for HER and OER performance.

Step 1: electron-sorption and proton-discharge

H2O + e + ZnCoFe2O4 ⇌ ZnCo-Hads·Fe2O4 +OH (Volmer step 120 mV dec−1)

Step 2: electro-desorption

ZnCo-Hads·Fe2O4 + H2O + e ⇌ ZnCoFe2O4 + H2 + OH (Heyrovsky step)

Step 3: chemical desorption

ZnCo-Hads·Fe2O4 + ZnCo-Hads Fe2O4 ⇌ 2ZnCoFe2O4+ H2 (Tafel reaction 40 mV dec−1)

The improved HER activity of ZnCoFe2O4, supported by its lower Tafel slope, indicates that the synergistic effect of zinc and cobalt enhances the availability of active sites and accelerates charge transfer, optimizing the reaction kinetics. Studies on ZnCoFe2O4 have demonstrated that the interaction between Zn and Co species facilitates the formation of unsaturated coordination sites and oxygen vacancies, significantly improving reactant adsorption and catalytic efficiency.90 Similarly, in perovskite oxides such as CaFe1−xCoxO3, the substitution of Co4+ into the Fe matrix has been reported to increase oxygen evolution reaction (OER) activity due to enhanced electronic interactions between Fe4+ and Co4+ ions.91 These findings align with our experimental observations, where the incorporation of Zn2+ and Co2+ into the spinel structure of ZnCoFe2O4 results in superior catalytic performance.

The integration of cobalt (Co) into zinc (Zn) ferrite remarkably enhances the HER kinetics by shrinking the energy barriers correlated with the reaction steps. This enhancement in catalytic performance is attributed to the synergistic effect of cobalt and zinc ions within the spinel lattice, which optimizes electron transfer and increases the availability of active sites. Its dual functionality can significantly reduce the overall cost of water-splitting systems, making it a promising candidate for sustainable hydrogen production.

Long-term stability is a critical parameter for the practical implementation of electrocatalysts in sustainable hydrogen production. To evaluate the durability of the ZnCoFe2O4 catalyst, a 14-hour chronoamperometric test was performed at a constant applied voltage of 0.6 V for the HER, as shown in Fig. 9(e). Remarkably, the catalyst retained 94% of its initial current density after 14 hours of continuous operation, demonstrating excellent stability and resilience under prolonged electrochemical conditions. This durability is primarily attributed to the enhanced ion penetration at the electrode–electrolyte interface, which occurs due to the reduction in ion diffusion resistance. The electrochemical activation of the ZnCoFe2O4 material during the testing process facilitates the generation of abundant active sites, thereby improving the interaction between the catalyst surface and reactant ions. These active sites significantly enhance the efficiency of the catalytic process by providing more accessible pathways for charge transfer.92,93 In addition, the catalyst's stability (73%) toward the oxygen evolution reaction (OER) was examined using a separate 14-hr chronoamperometric test at an applied potential of 1.6 V, as shown in Fig. 9(f), highlighting the durability of ZnCoF2O4 for both HER and OER processes.

In addition to HER performance, the ZnCoFe2O4 catalyst was also evaluated for overall water-splitting in a 6 M KOH solution. The tests demonstrated the catalyst's superior performance in both HER and OER processes, further confirming its bifunctional catalytic capabilities. This high efficiency can be attributed to the synergy between cobalt and zinc ions within the spinel lattice, which optimizes the electronic structure, enhances charge transfer properties, and increases the density of active sites.89

As shown in Fig. 10, cyclic voltammetry (CV) was conducted to evaluate the redox properties of mono- and binary ferrite electrocatalysts, ZnCoFe2O4, CoFe2O4, and ZnFe2O4, over a voltage range of 0.00 V to 0.60 V at scan rates (SRs) varying from 5 mV s−1 to 100 mV s−1. Fig. 10(a–d) shows the CV curves of these ferrites, highlighting their distinct redox behavior, particularly at a scan rate of 20 mV s−1. All samples exhibited redox peaks attributed to the OH and Fe3+/Fe4+ redox couple, essential for catalytic activity in HER and OER processes. Notably, ZnCoFe2O4 exhibits more pronounced redox peaks relative to CoFe2O4 and ZnFe2O4, suggesting enhancement in its catalytic efficiency. Such findings highlight the importance of faster electron transfers at the electrode interface for efficient hydrogen evolution and oxygen evolution processes.


image file: d5ta01180a-f10.tif
Fig. 10 (a) Cyclic voltammogram (CV) of ZnCoFe2O4 at varying scan rates (SRs) (5–100 mV s−1), showing distinct redox peaks associated with the OH and Fe3+/Fe4+ couple. (b) CV of CoFe2O4, under identical conditions, demonstrating lower redox activity compared to ZnCoFe2O4. (c) CV of ZnFe2O4, exhibiting weaker redox behavior relative to ZnCoFe2O4 and CoFe2O4. (d) Comparison of CV curves of ZnFe2O4, CoFe2O4, and ZnCoFe2O4 at a scan rate of 20 mV s−1, highlighting the superior redox performance of ZnCoFe2O4. (e) EIS Nyquist plots for ZnFe2O4, CoFe2O4, and ZnCoFe2O4, with the inset showing the high-frequency region. ZnCoFe2O4 exhibits the lowest charge transfer resistance, indicating superior charge transport and catalytic performance.

The ZnCoFe2O4 electrocatalyst displayed a well-defined oxidative peak at a higher potential compared to CoFe2O4 and ZnFe2O4, indicative of its superior redox reaction kinetics and higher electrochemical activity. This higher oxidative potential demonstrates the improved interaction of ZnCoFe2O4with the electrolyte, enhancing charge transfer at the electrode–electrolyte interface. The CV curves at different scan rates further reveal that ZnCoFe2O4 exhibits more pronounced redox features, suggesting a greater density of active sites available for catalytic reactions.

These findings underscore the superior performance of ZnCoFe2O4 as an electrocatalyst, demonstrating its effectiveness in facilitating both the HER and OER. The enhanced redox properties, combined with higher peak potentials and pronounced redox behavior, highlight its potential as a highly efficient and durable electrocatalyst for water-splitting applications. This performance improvement is attributed to the synergistic effect of cobalt and zinc substitution, which optimizes the electronic structure and increases active site availability, making ZnCoFe2O4 a promising candidate for sustainable energy applications.

Electrochemical impedance spectroscopy (EIS) was utilized to investigate the impedance response of ZnCoFe2O4, CoFe2O4, and ZnFe2O4 electrocatalysts under an applied alternating current (AC) signal across a range of frequencies. The Nyquist plots in Fig. 10(e) display the real impedance (Z′) on the x-axis and the imaginary impedance (Z′′) on the y-axis, offering insights into the intrinsic resistance of the electrode material and the interfacial resistance between the electrode and the electrolyte. This analysis evaluates the conductivity of the samples and the electron transport mechanisms at the electrode–electrolyte interface.94 A fitted equivalent circuit model was used to interpret the EIS data, incorporating elements such as the constant phase element (CPE) and Warburg impedance (W) to represent diffusion processes. The solution resistance (Rs) reflects the resistance offered by the electrolyte solution, which is critical for determining the system's overall impedance and electrochemical performance. Lower Rs values indicate improved ionic conductivity and reduced energy losses, facilitating better catalytic efficiency. The impedance behavior reflects minimal transfer limitation, and the dominant features are solution resistance (Rs) and Warburg diffusion (W). As summarized in Table 3, ZnCoFe2O4 exhibited a significantly lower Rs compared to CoFe2O4 and ZnFe2O4, indicating enhanced ionic transport and superior electrolyte conductivity.

Table 3 Solution resistance of ZnCoFe2O4, CoFe2O4, and ZnFe2O4a
Materials R s (Ω) W (s × s1/2)
a Furthermore, the Warburg impedance (W) decreased from 0.233 S s1/2 for ZnFe2O4 to 0.158 S s1/2 for ZnCoFe2O4, demonstrating more rapid diffusion of reactants and products within the electrolyte and improved mass transport to the electrode surface.
ZnCoFe2O4 1.58 158.9 × 10−3
CoFe2O4 1.83 200.7 × 10−3
ZnFe2O4 2.96 233.0 × 10−3


These EIS results are consistent with the enhanced OER and HER activities previously reported for ZnCoFe2O4, confirming its superior electrochemical performance. The combination of reduced solution resistance, lower charge transfer impedance, and faster diffusion kinetics highlights ZnCoFe2O4 as a promising material for improving water-splitting efficiency in practical applications. These findings underscore its potential for use in energy-efficient hydrogen production systems.

Conclusion

In this comparative study, we have successfully optimized the synthesis of the binary ferrite composite ZnCoFe2O4 and thoroughly characterized its structural, morphological, and elemental properties. Our findings highlight the synergistic enhancement of photocatalytic water splitting with the assistance of sound waves, emphasizing the crucial role of piezoelectric elements in piezo-photocatalysis. The improved HER activity, supported by a lower Tafel slope and enhanced charge transfer efficiency, underscores the significance of cobalt and zinc ions in optimizing reaction kinetics and increasing active site availability. Metal ferrites prove to be versatile catalysts for various water-splitting processes, with process selection contingent on the specific application. While both piezo-photocatalytic and electro-catalytic systems offer distinct advantages and limitations, ZnCoFe2O4 is suitable for both processes. Furthermore, when comparing MOFs and ferrites, the latter outperforms the former, with binary ferrite composites demonstrating superior efficiency over their mono-ferrite counterparts. This work underscores the promise of cobalt-substituted zinc ferrites in outperforming traditional HER and OER electrode materials in terms of overpotential and underwater stability and cost-effectiveness. However, practical implementation challenges remain, particularly regarding large-scale synthesis, long-term durability under industrial conditions, and potential deactivation due to structural changes over prolonged operation. Future research should focus on optimizing material stability, investigating alternative dopants to further enhance catalytic activity, and integrating these catalysts into scalable water-splitting systems.

Data availability

The data that supports the findings of this study are available from the corresponding authors upon reasonable request.

Author contributions

S. J., F. M., M. Z. K., S. J., J. H. K., & M. S. conceptualized the study. S. J., F. M., S. J., M. A., & M. S. characterized and analyzed. S. J., F. M., M. S., M. A. & A. M. did the writing and characterization. M. Z. K. & J. H. K. carried out the measurement. S. J., F. M., M. S., M. A. & A. M. conducted the electrochemical experiment. S. J., F. M., M. S., A. M., A. H. B., & M. B. K. N. commented on the manuscript. All the authors discussed the results and contributed to the manuscript preparation.

Conflicts of interest

The authors declare no conflict of interest.

Acknowledgements

This work was jointly supported by the National Research Program of Universities (NRPU), Higher Education Commission (HEC), Pakistan research fund “Development of lab/pilot scale facilities for the production of piezoelectric material for multilayer energy devices” 20-15673/NRPU/R&D/HEC/2021 and by the MSIT (Ministry of Science and ICT), Korea, under the ITRC (Information Technology Research Centre) support program (IITP-2025-RS-2020-II201655, 50%) supervised by the IITP (Institute of Information and Communications Technology Planning And Evaluation).

References

  1. L. Chen, et al., Separating hydrogen and oxygen evolution in alkaline water electrolysis using nickel hydroxide, Nat. Commun., 2016, 7, 11741 CrossRef CAS PubMed.
  2. M. Ali, et al., Current Trends in Nanoscale Interfacial Electrode Engineering for Sulfide-Based All-Solid-State Li-Ion Batteries, Energy Technol., 2021, 9(5), 2001096 CrossRef CAS.
  3. M. B. Hanif, et al., Innovative advances and challenges in solid oxide electrolysis cells: Exploring surface segregation dynamics in perovskite electrodes, Mater. Sci. Eng. R Rep., 2024, 161, 100864 CrossRef.
  4. H. S. Gujral, et al., Metal nitride-based nanostructures for electrochemical and photocatalytic hydrogen production, Sci. Technol. Adv. Mater., 2022, 23(1), 76–119 CrossRef CAS PubMed.
  5. J. Tang, et al., Emerging Energy Harvesting Technology for Electro/Photo-Catalytic Water Splitting Application, Catalysts, 2021, 11(1), 142 CrossRef CAS.
  6. S. C. Han, et al., Unraveling electrochemo-mechanical aspects of core–shell composite cathode for sulfide based all-solid-state batteries, J. Mater. Chem. A, 2024, 12(37), 24896–24905 RSC.
  7. L. Song, et al., Hydrogen production and mechanism from water splitting by metal-free organic polymers PVDF/PVDF-HFP under drive by vibrational energy, Fuel, 2022, 324, 124572 CrossRef CAS.
  8. P. Wu, et al., Electronic, Optical, piezoelectric properties and photocatalytic water splitting performance of Two-dimensional group IV-V compounds, Appl. Surf. Sci., 2023, 627, 157317 CrossRef CAS.
  9. X. Li, et al., Water Splitting: From Electrode to Green Energy System, Nano-Micro Lett., 2020, 12(1), 131 CrossRef CAS PubMed.
  10. U. Tariq, et al., Bridging the Gap between fundamentals and efficient devices: Advances in proton-conducting oxides for low-temperature solid oxide fuel cells, J. Power Sources, 2024, 613, 234910 CrossRef CAS.
  11. O. Chun, et al., Advances in low-temperature solid oxide fuel cells: An explanatory review, J. Power Sources, 2024, 610, 234719 CrossRef CAS.
  12. J. Eom, et al., Unveiling the dynamics of phase-transition from ferroelectric to relaxor behavior in Nd-doped BNT-based lead-free piezoelectric ceramics, J. Mater. Chem. C, 2024, 12, 19463–19475 RSC.
  13. Z. He, et al., Activating lattice oxygen in NiFe-based (oxy)hydroxide for water electrolysis, Nat. Commun., 2022, 13(1), 2191 CrossRef CAS PubMed.
  14. K. Jiang, et al., Nickel-cobalt nitride nanoneedle supported on nickel foam as an efficient electrocatalyst for hydrogen generation from ammonia electrolysis, Electrochim. Acta, 2022, 403, 139700 CrossRef CAS.
  15. A. Ursua, L. M. Gandia and P. Sanchis, Hydrogen Production From Water Electrolysis: Current Status and Future Trends, Proc. IEEE, 2012, 100(2), 410–426 CAS.
  16. X. Huang, et al., Insight into the piezo-photo coupling effect of PbTiO3/CdS composites for piezo-photocatalytic hydrogen production, Appl. Catal. B Environ., 2021, 282, 119586 CrossRef CAS.
  17. S. Tu, et al., Band shifting triggered ·OH evolution and charge separation enhanced H2O2 generation in piezo-photocatalysis of Silleń-Aurivillius-structured Bi4W0.5Ti0.5O8Cl, Chem. Eng. J., 2023, 465, 142777 CrossRef CAS.
  18. M. Ali, et al., Solvent exchange-induced facile recrystallisation and particle size control of sulphide solid electrolytes for all-solid-state Li-ion batteries, J. Mater. Chem. A, 2022, 10(48), 25471–25480 RSC.
  19. Q. Liu, et al., Na-Sm Bimetallic Regulation and Band Structure Engineering in CaBi2Nb2O9 to Enhance Piezo-photo-catalytic Performance, Adv. Funct. Mater., 2023, 33(38), 2303736 CrossRef CAS.
  20. W. Amdouni, et al., BiFeO3 Nanoparticles: The "Holy-Grail" of Piezo-Photocatalysts?, Adv. Mater., 2023, 35(31), e2301841 CrossRef PubMed.
  21. Y. Zhang, et al., High Efficiency Water Splitting using Ultrasound Coupled to a BaTiO3 Nanofluid, Adv. Sci. (Weinh.), 2022, 9(9), e2105248 Search PubMed.
  22. L. L. G. Al-Mahamad, Analytical study to determine the optical properties of gold nanoparticles in the visible solar spectrum, Heliyon, 2022, 8(7), e09966 CrossRef CAS PubMed.
  23. Z. Li, et al., Surface-Polarity-Induced Spatial Charge Separation Boosts Photocatalytic Overall Water Splitting on GaN Nanorod Arrays, Angew Chem. Int. Ed. Engl., 2020, 59(2), 935–942 CrossRef CAS PubMed.
  24. M. L. Xu, et al., Piezo-Photocatalytic Synergy in BiFeO3 @COF Z-Scheme Heterostructures for High-Efficiency Overall Water Splitting, Angew Chem. Int. Ed. Engl., 2022, 61(44), e202210700 CrossRef CAS PubMed.
  25. P. Jia, J. Li and H. Huang, Piezocatalysts and Piezo-Photocatalysts: From Material Design to Diverse Applications, Adv. Funct. Mater., 2024, 34(44), 2407309 CrossRef CAS.
  26. Y. Liu, et al., Internal-Field-Enhanced Charge Separation in a Single-Domain Ferroelectric PbTiO3 Photocatalyst, Adv. Mater., 2020, 32(7), 1906513 CrossRef CAS PubMed.
  27. H. Gulab, et al., Advancements in zinc oxide nanomaterials: synthesis, properties, and diverse applications, Nano-Struct. Nano-Objects, 2024, 39, 101271 CrossRef.
  28. Q. Liu, et al., A (Bi2O2)2+ layer as a significant carrier generator and transmission channel in CaBi2Nb2O9 platelets for enhanced piezo-photo-catalytic performance, Nano Energy, 2023, 108, 108252 CrossRef CAS.
  29. Y. Feng, et al., Engineering spherical lead zirconate titanate to explore the essence of piezo-catalysis, Nano Energy, 2017, 40, 481–486 CrossRef CAS.
  30. S. Tu, et al., Piezocatalysis and Piezo-Photocatalysis: Catalysts Classification and Modification Strategy, Reaction Mechanism, and Practical Application, Adv. Funct. Mater., 2020, 30(48), 2005158 CrossRef CAS.
  31. R. Su, et al., Nano-Ferroelectric for High Efficiency Overall Water Splitting under Ultrasonic Vibration, Angew Chem. Int. Ed. Engl., 2019, 58(42), 15076–15081 CrossRef CAS PubMed.
  32. L. Chen, et al., Strong piezocatalysis in barium titanate/carbon hybrid nanocomposites for dye wastewater decomposition, J. Colloid Interface Sci., 2021, 586, 758–765 CrossRef CAS PubMed.
  33. M. Ali, et al., Thermal stability analysis of nitrile additives in LiFSI for lithium-ion batteries: An accelerating rate calorimetry study, Heliyon, 2024, 10(9), e29397 CrossRef CAS PubMed.
  34. S. Park, et al., Microstructure-based digital twin thermo-electrochemical modeling of LIBs at the cell-to-module scale, Transportation, 2024, 22, 100370 Search PubMed.
  35. W. Wang, et al., Boosting Efficiency in Piezo-Photocatalysis Process Using Poled Ba0.7Sr0.3TiO3 Nanorod Arrays for Pollutant Degradation and Hydrogen Production, ACS Appl. Mater. Interfaces, 2024, 16(16), 20497–20509 CAS.
  36. Y. Wang, et al., Band Engineering of BiFeO3 Nanosheet for Boosting Hydrogen Evolution by Synergetic Piezo-photocatalysis, ACS Sustain. Chem. Eng., 2024, 12(6), 2300–2312 CrossRef CAS.
  37. B. Zhong, et al., Advances in ultrasound-assisted photocatalyst synthesis and piezo-photocatalysts, J. Mater. Chem. A, 2023, 11(42), 22608–22630 RSC.
  38. C. Li, et al., Harnessing ultrasound in photocatalysis: Synthesis and piezo-enhanced effect: A review, Ultrason. Sonochem., 2023, 99, 106584 CrossRef CAS PubMed.
  39. H. Cai, et al., Ultrathin Bi4O5I2 nanosheets as an integrated piezo-photocatalyst: Super visible-light piezo-photocatalysis and synergistic catalytic mechanism, Appl. Surf. Sci., 2023, 635, 157771 CrossRef CAS.
  40. H. Lv, et al., The insight into the critical role of photoexcitation in manipulating charge carrier migration in piezo-photocatalytic S-scheme heterojunction, Mater. Today Phys., 2023, 37, 101212 CrossRef CAS.
  41. L. Pan, et al., Advances in Piezo-Phototronic Effect Enhanced Photocatalysis and Photoelectrocatalysis, Adv. Energy Mater., 2020, 10(15), 2000214 CrossRef CAS.
  42. W. Chen, et al., Hierarchical porous bimetal-sulfide bi-functional nanocatalysts for hydrogen production by overall water electrolysis, J. Colloid Interface Sci., 2020, 560, 426–435 CrossRef CAS PubMed.
  43. G. Tanvir, et al., Study of ferroelectric and piezoelectric response of heat-treated surfactant-based BaTiO3 nanopowder for high energy capacitors, Mater. Sci. Eng., B, 2023, 287, 116100 CrossRef CAS.
  44. D.-J. Shin, et al., Fabrication and stability of base metal electrode (Ni) on a perovskite oxide co-fired multilayer piezoelectric device, J. Mater. Chem. C, 2021, 9(31), 10101–10111 RSC.
  45. S. Wang, A. Lu and C. J. Zhong, Hydrogen production from water electrolysis: role of catalysts, Nano Convergence, 2021, 8(1), 4 CrossRef CAS PubMed.
  46. K. Xiao, et al., Bimetallic sulfide interfaces: Promoting destabilization of water molecules for overall water splitting, J. Power Sources, 2021, 487, 229408 CrossRef CAS.
  47. J. T. Mefford, et al., Water electrolysis on La(1-x)Sr(x)CoO(3-delta) perovskite electrocatalysts, Nat. Commun., 2016, 7, 11053 CrossRef CAS PubMed.
  48. A. V. N, et al., Engineering nanostructured spinel ferrites by co-substitution for total water electrolysis by preferential exposure of metal cations on the surface, Sustain. Energy Fuels, 2020, 4(8), 3915–3925 RSC.
  49. M. L. Xu, et al., Piezo-Photocatalytic Synergy in BiFeO3@COF Z-Scheme Heterostructures for High-Efficiency Overall Water Splitting, Angew. Chem., Int. Ed., 2022, 61(44), e202210700 CrossRef CAS PubMed.
  50. C. Cao, et al., Semisacrificial Template Growth of Self-Supporting MOF Nanocomposite Electrode for Efficient Electrocatalytic Water Oxidation, Adv. Funct. Mater., 2019, 29(6), 1807418 CrossRef.
  51. D. Senthil Raja, et al., In situ grown bimetallic MOF-based composite as highly efficient bifunctional electrocatalyst for overall water splitting with ultrastability at high current densities, Adv. Energy Mater., 2018, 8(23), 1801065 CrossRef.
  52. J. Zhang, et al., Metal–organic-framework-based photocatalysts optimized by spatially separated cocatalysts for overall water splitting, Adv. Mater., 2020, 32(49), 2004747 CrossRef CAS PubMed.
  53. A. M. Fehr, et al., Integrated halide perovskite photoelectrochemical cells with solar-driven water-splitting efficiency of 20.8%, Nat. Commun., 2023, 14(1), 3797 CrossRef CAS PubMed.
  54. Y. Zheng, et al., Sulfur-doped g-C3N4/rGO porous nanosheets for highly efficient photocatalytic degradation of refractory contaminants, J. Mater. Sci. Technol., 2020, 41, 117–126 CrossRef CAS.
  55. L. Zhang, et al., Pyrolytic modification of heavy coal tar by multi-polymer blending: preparation of ordered carbonaceous mesophase, Polymers, 2024, 16(1), 161 CrossRef CAS PubMed.
  56. Y. Xue, et al., Enhanced photocatalytic performance of iron oxides@ HTCC fabricated from zinc extraction tailings for methylene blue degradation: Investigation of the photocatalytic mechanism, Int. J. Miner. Metall. Mater., 2023, 30(12), 2364–2374 CrossRef CAS.
  57. T. F. Qahtan, et al., State-of-the-art, challenges and prospects of heterogeneous tandem photocatalysis, Coord. Chem. Rev., 2023, 492, 215276 CrossRef CAS.
  58. M. Ghorbani, et al., Modified BiFeO3/rGO nanocomposite by controlled synthesis to enhance adsorption and visible-light photocatalytic activity, J. Mater. Res. Technol., 2023, 22, 1250–1267 CrossRef CAS.
  59. P. Lin, et al., Research progress in the preparation of electrospinning MOF nanofiber membranes and applications in the field of photocatalysis, Sep. Purif. Technol., 2025, 356, 129948 CrossRef CAS.
  60. F. Mumtaz, et al., Optimization of z-scheme Bi0.5Na0.5TiO3/RGO-Co3O4 composite catalyst for water splitting reaction through piezo-photocatalysis, Int. J. Hydrogen Energy, 2024, 78, 1468–1480 CrossRef CAS.
  61. T. O. Ogundipe, et al., Nickel-cobalt phosphide terephthalic acid nano-heterojunction as excellent bifunctional electrocatalyst for overall water splitting, Electrochim. Acta, 2022, 421, 140484 CrossRef CAS.
  62. B. Shabbir, et al., Facile synthesis of Er-MOF/Fe2O3 nanocomposite for oxygen evolution reaction, Mater. Chem. Phys., 2022, 292, 126861 CrossRef CAS.
  63. B. L. Choudhary, et al., Low temperature field dependent magnetic study of the Zn0.5Co0.5Fe2O4 nanoparticles, J. Magn. Magn. Mater., 2021, 536, 168102 CrossRef CAS.
  64. H.-C. Roth, et al., Influencing factors in the CO-precipitation process of superparamagnetic iron oxide nano particles: A model based study, J. Magn. Magn. Mater., 2015, 377, 81–89 CrossRef CAS.
  65. H. Yan, et al., Influences of different synthesis conditions on properties of Fe3O4 nanoparticles, Mater. Chem. Phys., 2009, 113(1), 46–52 CrossRef CAS.
  66. A. P. Herrera, et al., Synthesis and agglomeration of gold nanoparticles in reverse micelles, Nanotechnology, 2005, 16(7), S618 CrossRef PubMed.
  67. V. Shakeel, et al., Biocompatible gelatin-coated ferrite nanoparticles: A magnetic approach to advanced drug delivery, Saudi Pharm. J., 2024, 32(6), 102066 CrossRef CAS PubMed.
  68. M. Abbas, et al., Hierarchical porous spinel MFe2O4 (M= Fe, Zn, Ni and Co) nanoparticles: Facile synthesis approach and their superb stability and catalytic performance in Fischer-Tropsch synthesis, Int. J. Hydrogen Energy, 2020, 45(18), 10754–10763 CrossRef CAS.
  69. H. L. Andersen, et al., Crystalline and magnetic structure–property relationship in spinel ferrite nanoparticles, Nanoscale, 2018, 10(31), 14902–14914 RSC.
  70. M. Dhiman, et al., Morphology controlled hydrothermal synthesis and photocatalytic properties of ZnFe2O4 nanostructures, Ceram. Int., 2016, 42(11), 12594–12605 CrossRef CAS.
  71. I. S. Jurca, et al., Structural study of CoFe2/CoFe2O4 multilayers, Surf. Sci., 2003, 529(1), 215–222 CrossRef CAS.
  72. P. Yaseneva, M. Bowker and G. Hutchings, Structural and magnetic properties of Zn-substituted cobalt ferrites prepared by co-precipitation method, Phys. Chem. Chem. Phys., 2011, 13(41), 18609–18614 RSC.
  73. D. S. Nikam, et al., Cation distribution, structural, morphological and magnetic properties of Co 1− x Zn x Fe 2 O 4 (x= 0–1) nanoparticles, RSC Adv., 2015, 5(3), 2338–2345 RSC.
  74. C. Singh, et al., Synthesis of zinc substituted cobalt ferrites via reverse micelle technique involving in situ template formation: a study on their structural, magnetic, optical and catalytic properties, Mater. Chem. Phys., 2015, 156, 188–197 CrossRef CAS.
  75. J. P. Singh, et al., Microstructure, local electronic structure and optical behaviour of zinc ferrite thin films on glass substrate, R. Soc. Open Sci., 2018, 5(10), 181330 CrossRef CAS PubMed.
  76. M. A. Golsefidi, et al., Hydrothermal method for synthesizing ZnFe2O4 nanoparticles, photo-degradation of Rhodamine B by ZnFe2O4 and thermal stable PS-based nanocomposite, J. Mater. Sci.: Mater. Electron., 2016, 27(8), 8654–8660 CrossRef CAS.
  77. R. K. Singh, FTIR spectroscopy of natural fluorite from Ambadongar, Gujarat, J. Geol. Soc. India, 2013, 81(2), 215–218 CrossRef CAS.
  78. M. M. El-Masry, et al., Selective photocatalytic reduction of nitroarenes into amines based on cobalt/copper ferrite and cobalt-doped copper ferrite nano-photocatalyst, J. Mater. Sci.: Mater. Electron., 2021, 32(13), 18408–18424 CrossRef CAS.
  79. K. Hadjiivanov, Adsorption properties of Fe-containing dealuminated BEA zeolites as revealed by FTIR spectroscopy, Microporous Mesoporous Mater., 2010, 131(1–3), 1–12 CrossRef CAS.
  80. M. M. El-Masry, R. Ramadan and M. K. Ahmed, The effect of adding cobalt ferrite nanoparticles on the mechanical properties of epoxy resin, Results Mater., 2020, 8, 100160 CrossRef.
  81. P. A. Vinosha, et al., Review on recent advances of zinc substituted cobalt ferrite nanoparticles: Synthesis characterization and diverse applications, Ceram. Int., 2021, 47(8), 10512–10535 CrossRef CAS.
  82. V. F. Kiselev and O. V. Krylov, Adsorption and catalysis on transition metals and their oxides. Springer Science & Business Media. 2012Vol. 9.
  83. S. Maity, et al., High surface area NiCoP nanostructure as efficient water splitting electrocatalyst for the oxygen evolution reaction, Mater. Res. Bull., 2021, 140, 111312 CrossRef CAS.
  84. A. Syed, et al., Improved electrocatalytic efficiency of nitrogen-doped Nb2CTx MXene in basic electrolyte for overall water splitting, Int. J. Hydrogen Energy, 2024, 83, 39–50 CrossRef CAS.
  85. C. Liu, et al., Recent progress of Ni-based catalysts for methanol electrooxidation reaction in alkaline media, ASEM., 2023, 2(2), 100055 Search PubMed.
  86. B. Ross, S. Haussener and K. Brinkert, Impact of Gas Bubble Evolution Dynamics on Electrochemical Reaction Overpotentials in Water Electrolyser Systems, J. Phys. Chem., 2025, 129(9), 4383–4397 CAS.
  87. E. Fabbri and T. J. Schmidt, Oxygen Evolution Reaction—The Enigma in Water Electrolysis, ACS Catal., 2018, 8(10), 9765–9774 CrossRef CAS.
  88. C. Wu, et al., Recommended electrochemical measurement protocol for oxygen evolution reaction, DeCarbon, 2025, 8, 100108 CrossRef.
  89. H. Belhadj, et al., 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(46), 20129–20137 CrossRef CAS.
  90. J. Liu, et al., Beneficial Synergistic Intermetallic Effect in ZnCo2O4 for Enhancing the Limonene Oxidation Catalysis, Inorg. Chem., 2023, 62(45), 18750–18757 CrossRef CAS PubMed.
  91. I. Yamada, et al., Synergistic Effect between Fe4+ and Co4+ on Oxygen Evolution Reaction Catalysis for CaFe1−xCoxO3, Mater. Trans., 2023, 64(9), 2097–2104 CrossRef CAS.
  92. Z. Zhao, et al., Engineering active and robust alloy-based electrocatalyst by rapid Joule-heating toward ampere-level hydrogen evolution, Nat. Commun., 2024, 15(1), 7475 CrossRef CAS PubMed.
  93. B. Zhang, et al., A strongly coupled Ru–CrOx cluster–cluster heterostructure for efficient alkaline hydrogen electrocatalysis, Nat. Catal., 2024, 7(4), 441–451 CrossRef CAS.
  94. M. Van Haeverbeke, M. Stock and B. J. I. A. De Baets, Equivalent electrical circuits and their use across electrochemical impedance spectroscopy application domains, Ieee Access, 2022, 10, 51363–51379 Search PubMed.

Footnotes

Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d5ta01180a
Both authors have contributed equally to this work.

This journal is © The Royal Society of Chemistry 2025
Click here to see how this site uses Cookies. View our privacy policy here.