Amarnath
Pasupathi
a,
Praveen
Kandasamy
bc,
Ranjith Kumar
Dharman
d,
Sivakumar
Govindarajan
b,
Tae Hwan
Oh
d,
Min Wook
Lee
c and
Yugeswaran
Subramaniam
*a
aApplied Thermal Plasma Laboratory, Department of Physics, Pondicherry University, Puducherry 605014, India. E-mail: yugeswaran@pondiuni.ac.in
bCentre for Engineered Coatings, International Advanced Research Centre for Powder Metallurgy and New Materials, Balapur, Hyderabad 500 005, Telangana, India
cInstitute of Advanced Composite Materials, Korea Institute of Science and Technology, Chudong-ro, Bongdong-eup, Jeonbuk 55324, Republic of Korea
dSchool of Chemical Engineering, Yeungnam University, Gyeongsan 38541, South Korea
First published on 14th May 2025
Electrocatalytic water splitting is a promising technique for producing sustainable hydrogen, but its effectiveness depends on the development of cost-effective and high-performance electrodes. In this work, phase-pure high entropy oxide (HEO) (Ni, Fe, Co, Cu, Mn)3O4 nanostructured coating electrodes were fabricated using a solution precursor plasma spray coating technique under optimized conditions with two different molar concentrations (1 M and 2 M) of solution precursors. This process enables precise deposition of a porous catalyst coating on stainless steel substrates, with an average thickness of 30 micrometers. The as-deposited coating shows a spinel structure, and its degree of crystallinity increases with higher molar concentrations of the solution precursors. The HEO coating electrodes demonstrate excellent activity in alkaline media for both the hydrogen evolution reaction (HER) and the oxygen evolution reaction (OER), with low overpotentials of 129 mV and 220 mV, respectively, at a current density of 10 mA cm−2. A two-electrode device was fabricated, and the results reveal that the required overall potential to achieve a current density of 10 mA cm−2 is 1.47 V only. This work highlights the potential of solution precursor plasma spray coating as a versatile and scalable approach for producing phase-pure HEO-based water-splitting electrodes, paving the way for large-scale hydrogen generation in sustainable energy systems.
To address this critical research gap, efforts are focused on the development of materials composed of highly active, stable, and Earth-abundant non-noble metals, which could provide both cost-effectiveness and superior catalytic performance for water oxidation.4 Extensive investigations have been conducted on a variety of transition metal-based compounds, including metal oxides, carbides, nitrides, phosphides, borides, selenides, and sulphides, as potential substitutes for noble metals. These materials are being explored to reduce catalyst costs while simultaneously improving catalytic efficacy. Numerous studies have demonstrated that first-row transition metals, such as nickel (Ni), manganese (Mn), and iron (Fe), exhibit exceptional intrinsic electrochemical activity for water splitting.5 Particularly, spinel structure metal oxides comprising multiple transition metals demonstrate superior catalytic activity compared to single-metal and bimetallic spinel oxides. This enhanced performance is attributed to their distinctive electronic structure, characterized by densely packed O2− anions and transition metal cations occupying both tetrahedral and octahedral coordination sites.6
Recent studies emphasize the emerging class of multi-metal oxides, termed high entropy oxides (HEOs), which integrate five or more metal cations into a single-phase crystal structure, positioning them as promising catalysts for replacing noble-metal-based catalysts in water splitting. These materials exhibit exceptional properties, including tuneable electrical, optical, and mechanical characteristics, as well as enhanced structural stability, arising from intrinsic elemental interactions.7 In contrast to conventional oxide materials, HEOs are characterized by distinctive features such as high entropy, lattice distortion, sluggish kinetics, and cocktail effects, all of which contribute to their unique catalytic performance.8
Duan et al. synthesized HEO (Fe, Co, Ni, Cr, Mn)3O4 nanoparticles using a solvothermal approach and utilized them as an electrocatalyst for the oxygen evolution reaction. This catalyst demonstrates excellent electrocatalytic performance with an overpotential of 288 mV at a current density of 10 mA cm−2 and outstanding stability for 95 h in a 1 M KOH solution.9 Amarnath et al. synthesized spinel structured (Ni, Co, Cr, Mn, V)3O4 HEO nanoparticles through a thermal plasma route and used them as an electrode material for electrochemical water splitting in an alkaline medium. They demonstrated outstanding OER and HER activity with excellent stability.10 Numerous other methods have been adopted to produce HEO nanoparticles, including hydrothermal synthesis,11 solid-state synthesis,12 sol–gel,13 co-precipitation,14 flame spray pyrolysis,15 and nebulized spray pyrolysis.16 However, in the fabrication of HEO electrodes for electrochemical performance studies, binders and other additives are commonly employed to form an integrated microstructure that ensures stable operation. The inclusion of these additives, along with the coating interface between the electrode surface and the HEO, can undermine the catalytic activity of the material. Therefore, there is a potential need to develop innovative synthesis methods that enable the deposition of additive-free HEO coatings with a uniform phase and microstructure, while ensuring robust adhesion to the electrode surface.
Solution precursor plasma spraying (SPPS) is a versatile technique for producing nanostructured metal oxide coatings on different substrates.17 This novel process involves injecting a formulated precursor solution containing the desired metal cations into a hot plasma jet, where the solution undergoes various stages of transformation to form in situ metal oxide particles before melting to form splat deposits on the substrate. It is pertinent to note that the single step consolidated deposits exhibit a fine structured microstructure with desirable porosity and adhesive strength.18 The SPPS technique has the potential to produce both equilibrium and non-equilibrium phases in multi-component oxide systems by optimizing the plasma spray conditions and liquid injection parameters. To the best of the authors' knowledge, single-phase HEO coatings have not yet been achieved using the solution precursor plasma spray technique. Depositing a phase pure HEO coating using thermal plasma is a difficult operation due to the high temperature exposure and heterogeneous reaction environment of the plasma medium. When utilising multi-component oxide-forming precursors, exposure to a high-temperature plasma jet often leads to different rates of vaporisation of certain substances in the feedstock solutions. This process can produce significant phase segregation within the high-entropy oxide (HEO) structure, promoting secondary phase formation. Therefore, meticulous choice of plasma spraying parameters, such as input power, solution feed rate, carrier gas, injection position, substrate-to-gun distance, substrate temperature, scanning velocity, and number of scans, is crucial for the production of single-phase HEO coatings.
In this study, an attempt is being made for the first time to develop an additive-free HEO nanostructured electrocatalyst coating using the solution precursor plasma spray technique under optimal operating conditions. Following preliminary parametric optimization studies, a spinel-structured HEO (Ni, Fe, Co, Cu, Mn)3O4 coating was deposited onto a stainless steel (SS-304) substrate using two different molar concentrations of the feedstock precursor solution (1 M and 2 M). The as-sprayed HEO coatings were used as electrode materials, and their electrochemical water-splitting performance was studied in a 1 M KOH electrolyte solution. Linear sweep voltammetry (LSV), electrochemical impedance spectroscopy (EIS), and chronopotentiometry tests were conducted to evaluate the water-splitting performance of the HEO coating electrode.
| Parameters | Value |
|---|---|
| Plasma gun | 9 MB |
| Arc current (A) | 450 |
| Arc voltage (V) | 70 |
| Primary gas, Ar (lpm) @ 5.5 bar | 65 |
| Secondary gas, H2 (lpm) @ 3.5 bar | 4 |
| Solution feed rate (mL min−1) | 50–55 |
| Solution molar concentration (mole) | 1 & 2 |
| Atomizing gas, Air (psi) | 20 |
| Stand-off distance (mm) | 50 |
| Nozzle exit diameter (mm) | 5.5 |
The plasma spray torch was mounted on a six-axis robot arm and the substrates were fixed in the sample holder. The movement path of the spray torch and the standard raster scan pattern were controlled by using an ABB robot arm (Model: IRB4400, ABB, Sweden). An atomizer nozzle was positioned to inject the precursor solution along the radial direction perpendicular to the plasma jet. Atomised precursor droplets are subjected to swift transformations while traversing along the plasma plume, starting with evaporation of solvents, gelation, pyrolysis, sintering, and melting, and ending with the deposition of fine structured coatings on the substrate. The schematic representation of the solution precursor plasma spraying process with coating formation steps is illustrated in Fig. 1. The HEO coatings deposited at 1 M and 2 M concentrations were named HEO-A1 and HEO-A2, respectively.
![]() | ||
| Fig. 1 Schematic representation of the solution precursor plasma spray process detailing the coating formation steps. | ||
:
1 (relative to the total catalyst loading on the electrode surface) was mixed in a mortar. NMP solvent was then added, and the mixture was continuously stirred using a pestle. Subsequently, a precise volume of catalyst ink was applied to a 1 cm2 area of a graphite sheet to form the working electrode. All requisite electrochemical studies such as cyclic voltammetry (CV), linear sweep voltammetry (LSV), electrochemical impedance spectroscopy (EIS), and chronopotentiometry were systematically performed. Prior to the electrochemical studies, Ar gas was purged for 30 min in a 1 M KOH solution. LSV measurements were employed to acquire the polarisation curve in 1 M KOH solution at a sweep rate of 5 mV s−1. The expression for calculating the overpotential, reversible hydrogen electrode (RHE) conversion, Tafel slope and turnover frequency is provided in the ESI.†
m space group. The absence of discernible secondary peaks in the XRD pattern of the HEO coatings indicates that the chosen processing parameters are ideal for achieving phase-pure HEO coatings. It is observed that the phase composition of HEO coatings remains largely unchanged with variations in the molar concentration of the precursor solution. However, the increase in the precursor solution's mole concentration leads to a noticeable increase in peak intensity, suggesting improved crystallinity and density of the coating. As the molar concentration of the precursor solution increases, the amount of material available for deposition during the plasma spray process also increases, thereby promoting nucleation and crystallite growth within the resulting microstructure.19 The average crystallite sizes of the HEO-A1 and HEO-A2 coatings were determined using the Debye–Scherrer equation, yielding values of 6.5 nm and 8.1 nm, respectively. Additionally, the crystallinity of the HEO-A1 and HEO-A2 coatings was quantified using the following equation, resulting in values of 82% and 90%, respectively.20The observed differences in crystallinity can be attributed to variations in the metal cation concentration of the precursor solution, which influences the evaporation rates of the precursor droplets within the high-temperature plasma. The increased concentration of metal cations, for a given plasma heat flux, likely promotes enhanced nucleation and subsequent crystal growth. Furthermore, the HEO coating's surface roughness also indicates an interesting trend. The obtained results showed that the solution precursors with high concentrations (HEO-A2) had lower surface roughness (8 ± 2 μm), whereas those with low concentrations (HEO-A1) had higher surface roughness (14 ± 3 μm). It was reported that the high roughness on the surface implies superior surface area and porousness, resulting in increased adsorption and desorption capacity and thereby, improved water splitting performance.21
Fig. 3 shows the cross-sectional and surface microstructure of as-deposited HEO coatings with two different molar concentrations. The successive steps of atomization, evaporation, fragmentation, gelation, sintering, and melting, and finally, a fine structured coating were achieved with smaller splat deposits (few micron size splats).22 The cross-sectional features of coatings show the typical characteristic morphology of a solution-derived microstructure, containing fine splats and pores with a few vertical cracks. The presence of un-pyrolyzed precursor materials in the coating composition promotes vertical crack formation due to the thermal stresses generated during spraying.22 The coating shows a highly porous and discontinuous microstructure at a low concentration of the solution precursor, as shown in Fig. 3a and b. The higher mole concentration of the solution precursor can achieve a well-bonded and relatively dense microstructure as observed in Fig. 3d and e.23 The HEO-A1 and HEO-A2 coatings show a thickness of about ∼30 μm. The surface of the coatings exhibits a cauliflower-like morphology, which is a characteristic feature of the SPPS process-derived microstructure.18 It contains fine-structured spherical clusters evenly distributed through the surface as shown in Fig. 3c and f. The fine-structured particles in the range of a few tens of nanometres were agglomerated on the spherical clusters resulting in the formation of nano-structured coatings. The porous microstructure of HEO-A1 coating has a high specific surface area, which is expected to enhance overall performance.
![]() | ||
| Fig. 3 Cross-sectional and surface microstructure of as-deposited coatings: (a–c) HEO-A1 and (d–f) HEO-A2 coatings. | ||
TEM images were acquired to conduct a detailed analysis of the microstructure of the as-deposited coating. Fig. 4a and b present the TEM image and SAED pattern of the as-deposited HEO-A1 coating. The SAED pattern exhibits well-defined diffraction rings, which are characteristic of polycrystalline materials, confirming the polycrystalline nature of the coating. The observed diffraction rings correspond to the (331), (422), and (511) crystallographic planes, respectively. This diffraction pattern provides further evidence of the structural integrity and crystallographic nature of the HEO-A1 coating, demonstrating that the coating formed via the plasma spray process maintains a well-defined polycrystalline microstructure. The high-resolution transmission electron microscopy (HR-TEM) images (Fig. 4c and d) reveal well-defined lattice fringes, which suggest the crystalline nature of the material. These lattice fringes are indicative of well-ordered atomic arrangements within the coating microstructure. The measured interplanar distances between adjacent lattice planes are 0.32, 0.26, 0.42, and 0.21 nm, corresponding to the (220), (311), (111), and (400) crystallographic planes, respectively. These values are consistent with the expected d-spacing for spinel-type structures, providing further evidence of the crystalline nature of the coating. The HR-TEM lattice fringes and SAED pattern are in strong agreement with the XRD result. Collectively, these complementary techniques confirm the phase purity, crystallinity, and structural integrity of the as-deposited coating through the SPPS process, validating the successful formation of a single-phase spinel structure.
![]() | ||
| Fig. 4 TEM images (a), SAED pattern (b), and HR-TEM images (c and d) of the as-deposited HEO-A1 coating. | ||
XPS analysis was conducted to investigate the surface chemistry and oxidation states of the as-deposited HEO coatings. Fig. 5 confirms the presence of all constituent elements (Ni, Fe, Mn, Cu, and Co) in both HEO coatings. The fine scan of Co 2p XPS spectra (Fig. 5a) was deconvoluted into four peaks with the spin–orbit characteristics of Cr2+ and Cr3+. The peaks present at 780.8 (Co 2p3/2) and 795.4 eV (Co 2p1/2) belong to Co3+, whereas, the peaks at 783.1 (Co 2p3/2) and 797.3 eV (Co 2p1/2) belong to Co2+. The peak at 803.8 eV is the satellite peak of Co 2p1/2. Fig. 5b illustrates the two main spin–orbital lines in Fe 2p spectra, and the peaks centered at 713.3 and 724.3 eV are ascribed to Fe 2p3/2 and Fe 2p1/2, respectively. The XPS peaks at 712.3 and 723.4 eV are attributed to Fe2+, while the peaks at 714.3 and 726 eV belong to Fe3+. The peak at 733 eV corresponds to the satellite peak of Fe3+. The XPS spectra of Mn 2p are shown in Fig. 5c, the peaks centered at 642.5 and 653.4 eV were ascribed to Mn2+ and the peaks located at 644.4 and 655.3 eV belong to Mn3+. In the Ni 2p XPS spectra (Fig. 5d), the peaks at 854.9 and 872.1 eV are ascribed to Ni2+, whereas the peaks at 856.3 and 873.6 eV are associated with Ni3+. Furthermore, the peaks at 861.9 and 880.6 eV correspond to the Ni 2p satellite peaks. The Cu 2p XPS spectra (Fig. 5e) were deconvoluted into Cu1+ peaks at 933.5 and 953.2 eV, and Cu2+ peaks at 935.2 and 955.1 eV. Satellite peaks at 942.2 and 962.6 eV were attributed to Cu2+. The XPS spectra of O 1s (Fig. 5f) were deconvoluted into three peaks O1, O2 and O3, which correspond to the metal–oxygen bond (529.8 eV), oxygen vacancies (531.5 eV) and surface absorbed oxygen (533.2 eV), respectively. HEO-A2 coatings have a higher intensity than HEO-A1 coatings due to the higher metallic concentration of the precursor solution. The obtained result was in agreement with the XRD result. The XPS results reveal variations in the intensities of each element in the HEO-A1 and HEO-A2 coatings without any shifts in binding energy, confirming that compositional differences occurred without changes in oxidation states during the deposition process, corresponding to the mole concentration of the precursor solution.
![]() | ||
| Fig. 5 XPS spectra of as-deposited HEO coatings: (a) Co 2p, (b) Fe 2p, (c) Mn 2p, (d) Ni 2p, (e) Cu 2p, and (f) O 1s. | ||
Fig. 7a shows the LSV polarization curves of various electrocatalysts towards HER performance, including HEO-A1, HEO-A2, a bare substrate (SS 304), and a commercial Pt/C catalyst. Notably, HEO-A1 achieved a significantly lower overpotential of 129 mV at a current density of −10 mA cm−2, outperforming both HEO-A2 (185 mV) and the bare substrate (270 mV) as shown in Fig. 7b. This indicates that HEO-A1 demonstrates superior electrocatalytic activity for the hydrogen evolution reaction, highlighting its potential as an efficient electrocatalyst compared to the other materials tested. The Tafel slope is used to assess the electrocatalytic activity of different catalysts and explain the reaction kinetics. As shown in Fig. 7c, HEO-A1 demonstrated a smaller Tafel value of 43 mV dec−1 than HEO-A2 (76 mV dec−1), which indicates that HEO-A1 involves a faster kinetic reaction. The following equations represent the HER mechanism in an alkaline medium:
| H2O + M + e− → M–H* + OH− (Volmer) |
| H2O + e− + M–H* → H2 + M + OH− (Heyrovsky) |
| 2M–H* → H2 + 2M (Tafel) |
Initially, H2O is electrochemically adsorbed on the catalysts (M–H*), resulting in the production of OH− (Volmer path). In the second stage, hydrogen is formed and released either chemically (Tafel path) or electrochemically (Heyrovsky path). Consequently, HEO-A1 indicates the Volmer–Heyrovsky mechanism where the electrochemical desorption is the rate-determining step. To assess the intrinsic HER electrocatalytic activity, turnover frequency (TOF) was measured and the results are displayed in Fig. 7d. It is shown that the HEO-A1 electrocatalyst exhibited a TOF valus of 9.12 s−1, which is greater than that of HEO-A2 (4.64 s−1). The chronopotentiometry test of HEO-A1 for the HER is shown in Fig. 7e. It can be observed that the voltage remains stable at a current density of −10 mA cm−2 in strong alkaline electrolyte, even after 45 h. Furthermore, the LSV curve after 45 h (Fig. 7f) demonstrates that the changes in overpotential are minimal, thus indicating the excellent stability of the coating microstructure. Table S2† compares different HEO-based electrocatalysts for the HER in 1 M KOH solution.
| S. No. | Electrocatalyst | Substrate | Cell voltage (V) | Current density (mA cm−2) | Ref. |
|---|---|---|---|---|---|
| 1 | (Li, Fe, Co, Ni, Cu, Zn)O | Carbon paper | 1.89 | 30 | 24 |
| 2 | (Fe, Co, Ni, Cu, Zn)3O4 | Ni foam | 1.65 | 10 | 25 |
| 3 | (Ni, Co, Cr, Mn, V)3O4 | Ni foam | 1.596 | 10 | 10 |
| 4 | (Ni, Co, Cr, Mn, Mo)3O4 | Ni foam | 1.58 | 10 | 2 |
| 5 | (Ni, Fe, Co, Cu, Mn)3O4 | SS plate | 1.47 | 10 | This work |
Following a comprehensive electrochemical investigation, it was significant to explore the electrocatalyst's structural, morphological, and oxidation state tenacity. The XRD analysis was performed to assess the structural change in HEO-A1 after the stability test, as shown in Fig. S4.† The results obtained indicate that the XRD pattern remains largely unchanged, with only a slight shift in the diffraction peaks. During the stability test, the electrochemical process encounters numerous redox cycles, which results in crystal lattice expansion or contraction. This leads to peak shifts in the XRD pattern. This suggests that the overall crystal structure of the material is stable, with minimal alterations occurring under the tested conditions. After the stability test, the electrocatalyst's surface chemistry and oxidation state were analysed using XPS, as shown in Fig. S5.† The results reveal that all the constituent metal components were still present in the HEO coatings following the stability test. In contrast to additive-based catalysts, additive-free catalysts show strong adherence to the substrate, reducing mechanical failure, preventing chemical degradation, and improving electrical conductivity during long-term operation. Hence, the utilization of the additive-free SPPS coating deposition technique not only improves long-term stability but also provides a significant advantage in terms of electrochemical performance. Similar to the observations made in phase analysis, the XPS data also confirm that there was no significant change in the oxidation states of the constitution metals, although a marginal shift in the binding energies was observed. This suggests that the metal cations were directly or indirectly involved in the water-splitting reaction, with only minimal changes in their chemical state. The HEO-A1 coating was characterized after the stability test using a FE-SEM to observe any changes in surface morphology. The microstructural observation results, as shown in Fig. S6,† confirm that there were no changes in the surface morphology of the coating after the electrochemical test, and the as-sprayed cauliflower-like structure was retained. The aforementioned investigations demonstrate that the electrocatalytic efficacy of the water splitting reaction is enhanced by the synergistic effect of the constituent metal components in high-entropy oxides and the processing approach.
![]() | ||
| Fig. 9 LSV result of the (a) OER, and (b) HER of the as-synthesized HEO nanopowder coated electrode. | ||
Compared to the SPPS coated electrode, the HEO powder-coated electrode exhibits significantly lower electrocatalytic performance in water splitting, with approximately 18% lower activity for the OER and 419% lower activity for the HER. The reduced performance can be attributed to the presence of an additive in the dense coating and the formation of a large air gap between the coating and the electrode interface. This air gap increases the resistivity, hindering charge mobility across the interface during electrolysis and consequently impairing the overall electrocatalytic efficiency of the electrode. In contrast, the SPPS-produced additive-free coating exhibits a microstructure with considerable porosity and vertical cracks. This distinctive microstructure facilitates the clear passage of the electrolyte flow through most of the coating microstructure, thus enriching charge transport and promoting more efficient electrolysis. Moreover, the coating formation leads to strong mechanical bonding with the substrate, which significantly reduces the interface resistance in the electrocatalytic-current collector. This improvement in interface conductivity enhances charge mobility and contributes to greater durability during electrochemical cycling. Additionally, the rough surface of the coating with a higher surface area-to-volume ratio and its inherent porosity, play a crucial role in facilitating bubble formation during the evolution of oxygen and hydrogen. The porous nature of the coating provides nucleation sites for gas bubbles, which can improve the overall efficiency of both reactions by promoting effective gas release and minimizing bubble accumulation at the electrode surface.26 These factors collectively contribute to enhancing the water-splitting performance of the coating microstructure.
Thermodynamic calculations indicate that a potential of 1.23 V is required to split water molecules under ideal conditions.27 However, achieving this thermodynamic limit through electrochemical processes presents significant challenges due to the intrinsic and structural properties of electrode materials, as well as the complexities associated with processing conditions. Currently, a potential of approximately 1.4 V has been attained for water splitting through electrochemical methods, utilizing noble metals such as platinum, iridium, and ruthenium, in addition to materials based on selenides, carbides, phosphides, and nitrides.28 These materials have been engineered with various structural and morphological modifications to enhance their catalytic efficiency, reduce overpotentials, and optimize their electrochemical performance under typical operational conditions.
Interestingly, the present study identifies a highly promising water-splitting electrode material composed of transition metals (Ni, Fe, Co, Cu, and Mn) arranged in a spinel structure within a high-entropy oxide configuration. This additive-free and structurally unique microstructure was synthesized via the SPPS process. Notably, the specific microstructural features of the HEO, such as its porosity, rough surface, and mechanical bonding with the substrate, significantly reduce the interface resistance and improve charge mobility. As a result, this material requires only a modest overpotential of 1.47 V to achieve a current density of 10 mA cm−2 for water splitting, demonstrating its highly efficient catalytic activity, positioning HEO-based materials prepared through the SPPS coating process as a promising alternative for scalable and efficient water-splitting applications.
Footnote |
| † Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d5se00479a |
| This journal is © The Royal Society of Chemistry 2025 |