High-entropy Cr(NiFeCoV)2O4 catalysts via CO2 laser thermal shock: advancing electrochemical water oxidation with multi-metal synergy

Sharanya Kannan Anbarasu a, Raja Arumugam Senthil a, Sieon Jung a, Anuj Kumar b, Mohd Ubaidullah c and Myong Yong Choi *ad
aDepartment of Chemistry (BK21 FOUR), Research Institute of Advanced Chemistry, Gyeongsang National University, Jinju 52828, Republic of Korea. E-mail: mychoi@gnu.ac.kr
bNano-Technology Research Laboratory, Department of Chemistry, GLA University, Mathura, Uttar Pradesh 281406, India
cDepartment of Chemistry, College of Science, King Saud University, Riyadh 11451, Saudi Arabia
dCore-Facility Center for Photochemistry & Nanomaterials, Gyeongsang National University, Jinju 52828, Republic of Korea

Received 25th February 2025 , Accepted 9th May 2025

First published on 9th May 2025


Abstract

Electrocatalytic overall water splitting (OWS) is a promising technology for sustainable hydrogen (H2) production. However, its practical application is hindered by the sluggish kinetics of the anodic oxygen evolution reaction (OER). To address this challenge, high-entropy oxides have emerged as promising OER electrocatalysts owing to their tunable composition and synergistic effects among constituent elements. In this study, we present the fabrication of a spinel-structured high-entropy Cr(NiFeCoV)2O4 (HE-Cr(NiFeCoV)2O4) catalyst using a rapid continuous-wave CO2 laser thermal-shock method. The resulting HE-Cr(NiFeCoV)2O4 catalyst demonstrated excellent electrochemical OER performance in 1 M KOH electrolyte, achieving a low overpotential of 284 mV at 10 mA cm−2 and maintaining long-term stability over 100 h at 50 mA cm−2. Furthermore, an OWS electrolyzer assembled with HE-Cr(NiFeCoV)2O4 as the anode and Pt/C as the cathode operated at a low cell voltage of 1.57 V at 10 mA cm−2 for efficient H2 production. In situ Raman spectroscopy confirmed the surface formation of active FeOOH species during OER, while density functional theory calculations revealed how the multi-metal synergy within a single lattice modulated the electronic structure, thereby enhancing the OER activity of Cr(NiFeCoV)2O4. This study establishes a cost-effective and energy-efficient pathway for developing advanced multicomponent electrocatalysts for clean energy applications.


1. Introduction

The increasing global energy demands and the adverse effects of climate change driven by reliance on nonrenewable fossil fuels underscore the urgent need for sustainable energy solutions.1,2 In this context, renewable hydrogen (H2) has emerged as a promising clean energy carrier owing to its high energy density, environmental compatibility, and zero carbon emissions.3,4 Among various H2 production technologies, water electrolysis is widely recognized as an efficient and environmentally friendly process that facilitates overall water splitting (OWS) powered by renewable energy sources.5,6 This process comprises the hydrogen evolution reaction at the cathode and the oxygen evolution reaction (OER) at the anode, yielding H2 and O2 gases, respectively.7,8 However, the large-scale implementation of OWS for H2 production is hindered by the slow kinetics and high overpotential of the anodic OER, necessitating the development of efficient and robust electrocatalysts to minimize energy input and enhance overall efficiency of the OWS. Although noble metal-based catalysts such as IrO2 and RuO2 exhibit exceptional electrochemical OER activity, their scarcity, high cost, and limited durability restrict their widespread application.9,10 This has prompted the search for low-cost, durable, and efficient alternatives for OER electrocatalysts.

Transition metals and their compounds, including oxides, sulfides, phosphides, carbides, and hydroxides, have been extensively studied as potential OER electrocatalysts.11–13 Among these, transition metal oxides-particularly those with rock-salt, perovskite, and spinel crystalline structures-have demonstrated exceptional electrocatalytic performance owing to their strong OH adsorption and surface-active *OOH intermediate generation capabilities via metal–oxygen (M–O) bonding, which enhances OER kinetics.14,15 Spinel-structured metal oxides particularly stand out owing to their straightforward synthesis, structural and compositional versatility, mixed metal valence states (+2 and +3), excellent catalytic redox properties, and stability in strongly alkaline environments.16,17 Recently, high-entropy oxides (HEOs), comprising five or more equimolar cations alongside oxygen anions, have emerged as a transformative class of electrocatalysts, offering viable alternatives to noble metal catalysts in energy conversion and storage applications.18–20 The incorporation of multiple metal cations in a single HEO structure facilitates extensive electron transfer between metal cations, resulting in strong electron redistribution and abundant active sites for electrochemical reactions.21,22 Additionally, the lattice distortion induced by multiple metal species generates efficient surface-oxygen vacancies, enhancing electron redistribution and creating unsaturated coordination environments in the HEOs.23–25 Remarkably, the high-entropy effect in HEOs, induced by multiple metal components, increases configurational entropy and reduces Gibbs free energy (ΔG↓ = ΔHTΔS↑), thereby improving catalytic activity, structural stability, and synergistic interactions among constituent metals.26–28 These unique features endow HEO-based OER electrocatalysts with exceptional electrochemical performance and long-term stability, making them promising candidates for advancing OWS efficiency.

Encouraged by the superior advantages of HEO catalysts in OER, we aimed to develop a spinel-structured high-entropy Cr(NiFeCoV)2O4 (denoted as HE-Cr(NiFeCoV)2O4) electrocatalyst. The integration of Cr, Ni, Fe, Co, and V cations within a single oxide matrix is anticipated to enhance electrocatalytic activity for OER. Additionally, the synthesis method plays a decisive role in determining the production cost, purity, and catalytic performance of nanomaterials. Conventional furnace-based heating methods are frequently utilized for the synthesis of metal oxides, including HEO nanomaterials.29,30 However, these conventional approaches often require extended heating durations, substantial energy input, and may generate impurities as byproducts, thereby constraining their practical applicability. To mitigate these challenges, thermal-shock synthesis employing a continuous-wave CO2 laser has been introduced, facilitating rapid, localized heating that allows for precise control over phase formation, crystallinity, and defect engineering.31–33 Furthermore, this innovative technique reduces thermal gradients, prevents undesired phase transitions, and drastically reduces synthesis time and energy consumption. These benefits position thermal-shock synthesis as a versatile and efficient method for the fabrication of advanced nanomaterials, offering a viable alternative to traditional synthesis techniques.

Herein, we successfully synthesized a spinel-structured HE-Cr(NiFeCoV)2O4 catalyst for the first time via a rapid thermal-shock approach using continuous-wave CO2 laser technology and evaluated its OER performance. The HE-Cr(NiFeCoV)2O4 catalyst demonstrated outstanding OER performance, achieving a low overpotential of 284 mV at 10 mA cm−2 and long-term stability over 100 h at 50 mA cm−2 in 1 M KOH electrolyte. Furthermore, when integrated as the anode in an OWS electrolyzer with Pt/C as the cathode, the system required a low cell voltage of only 1.57 V at 10 mA cm−2 for H2 production and maintained long-term operational stability over 100 h at 50 mA cm−2, highlighting its practical applicability. To elucidate the origins of the superior performance of the HE-Cr(NiFeCoV)2O4 catalyst, advanced in situ Raman spectroscopy and theoretical density functional theory (DFT) simulations were employed. In situ Raman spectroscopy revealed the formation of surface-active FeOOH species on HE-Cr(NiFeCoV)2O4 during OER, providing real-time insights into catalyst surface transformations. Concurrently, DFT simulations showed strong electronic modulations within the HE-Cr(NiFeCoV)2O4 structure and lower energy barriers for reaction intermediates, contributing to its exceptional OER performance. Overall, this study integrates advanced synthesis techniques and in-depth structural analyses to pave the way for the development of efficient, cost-effective, and durable multicomponent electrocatalysts for renewable energy applications.

2. Experimental section

2.1. Chemicals and reagents

Nickel chloride tetrahydrate (NiCl2·2H2O, 96%) was obtained from Yakuri Chemicals, South Korea. Iron chloride tetrahydrate (FeCl3·6H2O, 98%), cobalt chloride tetrahydrate (CoCl2·6H2O, 98.5%), chromium chloride tetrahydrate (CrCl3·6H2O, 98%), vanadium chloride (VCl3·4H2O, 99%), absolute ethanol (C2H5OH, HPLC grade, 99.9%), potassium hydroxide (KOH, extra pure flakes, 93%), and commercial iridium oxide (IrO2) were sourced from Daejung Chemicals, South Korea. All chemicals used were of analytical-grade quality and were employed without further purification.

2.2. Synthesis of the HE-Cr(NiFeCoV)2O4 catalyst

The spinel-structured HE-Cr(NiFeCoV)2O4 catalyst was synthesized using a novel CO2 laser-induced rapid thermal-shock technique. The synthesis began with the preparation of metal hydroxide precursors via coprecipitation. Specifically, 0.2 mmol of CrCl3·6H2O, 0.1 mmol of NiCl2·2H2O, 0.1 mmol of FeCl3·6H2O, 0.1 mmol of CoCl2·6H2O, and 0.1 mmol of VCl3 were dissolved in 20 mL of deionized water in a beaker under magnetic stirring for 30 min. Next, 20 mL of 0.1 M NaOH solution was added to the reaction mixture, followed by stirring for additional 3 h to ensure complete precipitation. The resulting metal hydroxide precipitates were collected by centrifugation, thoroughly washed with deionized water and absolute ethanol, and dried at 80 °C for 3 h. Subsequently, 5 mg of the dried metal hydroxide precursors was placed in an alumina crucible and subjected to CO2 laser irradiation at 14 W power using a wide-lens configuration to ensure uniform exposure across the sample surface. This thermal-shock treatment, lasting 2 min, caused the metal hydroxides to decompose and form the spinel-structured HE-Cr(NiFeCoV)2O4. The resultant HE-Cr(NiFeCoV)2O4 product was washed with absolute ethanol to remove surface moisture and dried at 80 °C for 3 h before characterization and experimentation. For comparison, a bimetallic spinel-structured CrCo2O4 catalyst was synthesized using 0.2 mmol of CrCl3·6H2O and 0.4 mmol of CoCl2·6H2O during the coprecipitation step, while all other procedures remained identical. This comparison aims to investigate the influence of incorporating multiple metals (Ni, Fe, Co, and V) HE-Cr(NiFeCoV)2O4, as opposed to using only Co in CrCo2O4, on the OER performance.

Details on material characterization techniques, electrochemical assessments, in situ Raman spectroscopy, and computational methods are provided in the ESI.

3. Results and discussion

3.1. Material synthesis and physicochemical characterizations

Fig. 1a illustrates the schematic of the two-step synthesis procedure for the HE-Cr(NiFeCoV)2O4 catalyst. The synthesis involves preparing metal hydroxide precursors via simple coprecipitation, followed by their transformation into HE-Cr(NiFeCoV)2O4 through rapid thermal-shock treatment using continuous-wave CO2 laser technology. This innovative method generates high-temperature thermal shocks, enabling the rapid phase transformation of metal hydroxides into corresponding metal oxides within a very short time-period. The continuous-wave CO2 laser technology presents a promising alternative to conventional, time-consuming furnace-heating methods for synthesizing various nanomaterials for diverse applications.
image file: d5ta01573a-f1.tif
Fig. 1 (a) Schematic of the synthesis process, (b–d) HR-TEM images, (e) SAED pattern, and (f) TEM-EDS mapping images of the HE-Cr(NiFeCoV)2O4 catalyst.

The surface morphologies and elemental compositions of the synthesized HE-Cr(NiFeCoV)2O4 were analyzed using field-emission scanning electron microscopy (FE-SEM), high-resolution transmission electron microscopy (HR-TEM), and energy-dispersive X-ray spectroscopy (EDS). FE-SEM (Fig. S1a and b) images revealed agglomerated nanoparticles in both CrCo2O4 and HE-Cr(NiFeCoV)2O4 catalysts. SEM-EDS mapping confirmed the uniform distribution of Cr, Ni, Co, Fe, V, and O in the HE-Cr(NiFeCoV)2O4 structure. The SEM-EDS spectra (Fig. S2) showed the elemental composition of CrCo2O4: 18.93 wt% Cr, 43.48 wt% Co, and 37.59 wt% O. In contrast, HE-Cr(NiFeCoV)2O4 catalyst showed elemental compositions of 19.75 wt% Cr, 12.53 wt% Ni, 11.85 wt% Fe, 12.92 wt% Co, 7.30 wt% V, and 35.65 wt% O, confirming the successful incorporation of multiple metals and its high-entropy nature. HR-TEM images of the HE-Cr(NiFeCoV)2O4 catalyst (Fig. 1b–d) further demonstrated agglomerated nanoparticles with well-defined lattice fringes, showing an interplanar spacing of 0.18 nm corresponding to the (311) crystalline plane of spinel CrCo2O4 structure. The selected-area electron diffraction (SAED) pattern (Fig. 1e) confirmed the polycrystalline nature of HE-Cr(NiFeCoV)2O4, with distinct crystal planes matching the spinel CrCo2O4 framework. Furthermore, TEM-EDS spectra (Fig. S3) and mapping (Fig. 1f) validated the uniform distribution of Cr, Ni, Fe, Co, V, and O with compositions of 21.45 wt%, 12.75 wt%, 11.33 wt%, 13.04 wt%, 7.06 wt%, and 34.37 wt%, respectively, closely aligning with SEM-EDS results. Based on these compositions, the total mixing entropy (ΔSmix) was calculated using the equation: image file: d5ta01573a-t1.tif, where R is the gas constant, Ci represents the mole fraction of the ith element, and n denotes the total number of components. The calculated ΔSmix value for HE-Cr(NiFeCoV)2O4 was 1.65R, further emphasizing its high-entropy characteristics.

To verify the high-entropy formation and phase purity of the HE-Cr(NiFeCoV)2O4 catalyst, X-ray diffraction (XRD) analysis was performed. The XRD patterns (Fig. 2a) of both CrCo2O4 and HE-Cr(NiFeCoV)2O4 catalysts exhibited similar diffraction peaks, consistent with cubic spinel oxide structure (PDF#22-1084).34,35 Notably, HE-Cr(NiFeCoV)2O4 showed peak shifts toward higher 2θ values compared to CrCo2O4, indicating lattice distortion caused by the incorporation of additional Ni, Fe, and V cations into the CrCo2O4 structure. This lattice distortion likely modulates the electronic properties of the HE-Cr(NiFeCoV)2O4 catalyst, enhancing its OER catalytic activity. These findings confirm the successful synthesis of a single-phase HE-Cr(NiFeCoV)2O4 catalyst with multiple metal cations and no additional phases, highlighting its excellent phase purity. Further structural characterization was performed using Raman spectroscopy, which revealed characteristic peaks for the CrCo2O4 catalyst (Fig. 2b) at 188, 459, 539, 592, and 660 cm−1 corresponding to the typical F2g, Eg, F2g, F2g, and A1g vibrational modes of cubic spinel oxides, respectively.17,36 These characteristic peaks shifted to higher wavenumbers in the HE-Cr(NiFeCoV)2O4 catalyst, attributable to modified vibrational modes resulting from enhanced M–O interactions and lattice distortions induced by the integration of multiple metals into a single spinel oxide structure. In addition, Fourier transform infrared spectra (Fig. S4) revealed distinct vibrational peaks at 487 and 632 cm−1 in CrCo2O4, corresponding to M–O stretching vibrations at octahedral and tetrahedral sites, respectively. These peaks were noticeably reduced in intensity for the HE-Cr(NiFeCoV)2O4 catalyst, likely owing to lattice distortions from Ni, Fe, and V cation incorporation into the CrCo2O4 framework. Brunauer–Emmett–Teller (BET) surface area analysis was performed to assess the influence of multiple metal components on the surface area of the HE-Cr(NiFeCoV)2O4 catalyst (Fig. 2c), exhibiting considerably higher nitrogen adsorption and desorption capacities, resulting in a higher BET surface area of 116.11 m2 g−1 compared to CrCo2O4 (85.37 m2 g−1). This increased surface area provides abundant surface-active sites and enhances surface adsorption of reaction intermediates, contributing to improved catalyst's OER performance.


image file: d5ta01573a-f2.tif
Fig. 2 (a) XRD patterns, (b) Raman spectra, and (c) BET N2 adsorption–desorption isotherms of CrCo2O4 and HE-Cr(NiFeCoV)2O4 catalysts. (d–i) Core-level XPS plots of (d) Cr 2p, (d) Ni 2p, (f) Fe 2p, (g) Co 2p, (h) V 2p, and (i) O 1s for the HE-Cr(NiFeCoV)2O4 catalyst.

Advanced X-ray photoelectron spectroscopy (XPS) was employed to analyze the chemical composition and oxidation states of the CrCo2O4 and HE-Cr(NiFeCoV)2O4 catalysts. The core-level XPS plots of the HE-Cr(NiFeCoV)2O4 catalyst are shown in Fig. S5 and 2d–i, while those of the CrCo2O4 catalyst are shown in Fig. S6a–d. The wide-range XPS plots confirmed the presence of Cr, Ni, Co, Fe, V, and O in HE-Cr(NiFeCoV)2O4 (Fig. S5), whereas only Cr, Co, and O were detected in CrCo2O4 (Fig. S6a), verifying the successful incorporation of multiple metal components into the HE-Cr(NiFeCoV)2O4 structure. The Cr 2p XPS plots of both catalysts revealed two doublet peaks at ∼575.8/585.5 and 578.1/587.3 eV (Fig. 2d and S6b), corresponding to Cr 2p3/2/2p1/2 spins associated with Cr3+ and Cr4+, respectively.37 The Ni 2p XPS plots of HE-Cr(NiFeCoV)2O4 splits into two doublet peaks at ∼854.8/872.4 and 857.6/875.1 eV (Fig. 2e), along with satellite peaks, corresponding to Ni 2p3/2/2p1/2 spins of Ni2+ and Ni3+, respectively.35 Similarly, the Fe 2p XPS plots of HE-Cr(NiFeCoV)2O4 (Fig. 2f) exhibited two doublet peaks at ∼709.9/712.3 and 723.2/725.9 eV along with satellite peaks, ascribed to Fe 2p3/2/2p1/2 spins of Fe2+ and Fe3+, respectively.38 The Co 2p XPS plots of both catalysts (Fig. 2g and S6c) showed two doublet peaks at ∼780.0/795.6 and 782.6/797.5 eV, with satellite peaks, corresponding to the Co 2p3/2/2p1/2 spins of Co3+ and Co2+, respectively. The V 2p XPS plots of HE-Cr(NiFeCoV)2O4 (Fig. 2h) showed two peaks at 516.2 and 523.6 eV, corresponding to the V 2p3/2 and V 2p1/2 spins of V3+, respectively.39,40 The O 1s XPS plots of both catalysts (Fig. 2i and S6d) exhibited a peak at ∼529.4 eV, corresponding to the lattice M–O bonds (where M = Cr/Co in CrCo2O4 or M = Cr/Ni/Fe/Co/V in HE-Cr(NiFeCoV)2O4) while additional peaks at 531.2 and 532.3 eV are ascribed to surface-adsorbed O2 and H2O molecules, respectively.36 These results confirmed the presence of Ni, Fe, and Co in mixed +2 and +3 oxidation states, Cr in mixed +3 and +4 oxidation states, and V in the +3 oxidation state in HE-Cr(NiFeCoV)2O4. The incorporation of multiple cations with mixed oxidation states substantially modulates the electronic structure, enhancing the intrinsic electrocatalytic activity of HE-Cr(NiFeCoV)2O4. In addition, ultraviolet photoelectron spectra of the CrCo2O4 and HE-Cr(NiFeCoV)2O4 (Fig. S7a and b) revealed Ecut-off values of 7.26 and 8.68 eV, respectively, where Ecut-off represents the kinetic energy cutoff of secondary electrons. The work function for HE-Cr(NiFeCoV)2O4 was calculated as 12.54 eV, compared to 13.96 eV for CrCo2O4, reflecting its modulated electronic structure, enhanced electron transport, and improved electrical conductivity.41 Reflection electron energy loss spectra (Fig. S8) revealed a prominent decrease in elastic peak intensity for HE-Cr(NiFeCoV)2O4 compared to CrCo2O4, highlighting the enhanced surface and electronic properties of the HE-Cr(NiFeCoV)2O4 catalyst—potentially accelerating its electrocatalytic activity for OER.42

3.2. Electrocatalytic performance of HE-Cr(NiFeCoV)2O4 toward OER

The electrocatalytic OER performance of the synthesized HE-Cr(NiFeCoV)2O4 catalyst was evaluated using a three-electrolyte configuration in 1 KOH electrolyte (Fig. 3a). For comparison, the OER performance of CrCo2O4 and commercial IrO2 catalysts was also assessed. As shown in Fig. 3b, the linear sweep voltammetry (LSV) plots recorded at 5 mV s−1 reveal that HE-Cr(NiFeCoV)2O4 exhibited considerably enhanced OER performance, as characterized by a greater negative shift in the onset potential and higher current densities compared to CrCo2O4 and IrO2 catalysts. Specifically, Fig. 3c indicates that HE-Cr(NiFeCoV)2O4 achieved a lower overpotential of 284 mV at mA cm−2, compared to the bare nickel foam (NF) substrate (440 V), CrCo2O4 (404 V), and IrO2 (375 V). This reduced overpotential underscores the superior OER activity of HE-Cr(NiFeCoV)2O4, attributed to its optimized electronic structure and abundant active sites resulting from the synergistic contributions of multiple metal cations.38 To further validate the OER kinetics of HE-Cr(NiFeCoV)2O4, Tafel slope analysis was conducted.43 As depicted in Fig. 3d, HE-Cr(NiFeCoV)2O4 exhibited a considerably smaller Tafel slope of 99 mV dec−1, compared to the bare NF substrate (243 mV dec−1), CrCo2O4 (186 mV dec−1), and commercial IrO2 (132 mV dec−1). This lower Tafel slope indicates an optimized reaction pathway with faster charge transfer kinetics, leading to increased intrinsic catalytic activity and lower energy barriers for reaction intermediates in HE-Cr(NiFeCoV)2O4 compared to CrCo2O4, thereby accelerating the OER process.23 In addition, the intrinsic catalytic activity of HE-Cr(NiFeCoV)2O4 was further assessed through mass activity and turnover frequency (TOF) analyses. Fig. S9 shows that HE-Cr(NiFeCoV)2O4 possessed the highest mass activity of 79.78 A g−1 at 1.7 V vs. reversible hydrogen electrode (RHE) among the tested catalysts. Similarly, the TOF values derived from cyclic voltammetry (CV) measurements (Fig. S10a–d) demonstrated that HE-Cr(NiFeCoV)2O4 exhibited the highest TOF of 0.71 s−1 at 1.7 V vs. RHE among the tested catalysts (Fig. 3e). These findings highlight the role of multi-metal environment in enhancing the utilization of active sites and intrinsic catalytic activity of HE-Cr(NiFeCoV)2O4, thereby boosting OER performance.13 The charge transfer kinetics of HE-Cr(NiFeCoV)2O4 during OER were analyzed using electrochemical impedance spectroscopy.38 Nyquist plots recorded at 1.5 V vs. RHE (Fig. 3f) show that HE-Cr(NiFeCoV)2O4 exhibited a charge transfer resistance (Rct) of 0.75 Ω—notably lower compared to CrCo2O4 (6.26 Ω) and commercial IrO2 (3.87 Ω). This reduced Rct manifests the efficient charge transfer abilities of HE-Cr(NiFeCoV)2O4, attributed to its enhanced electrical conductivity and optimized electronic structure, which are facilitated by the synergistic interaction within its multicomponent composition, ultimately resulting in superior OER performance.44
image file: d5ta01573a-f3.tif
Fig. 3 Electrocatalytic OER performance of the HE-Cr(NiFeCoV)2O4 catalyst in 1 M KOH electrolyte: (a) schematic of the three-electrode setup, where WE, CE, and RE denote the working, counter, and reference electrodes, respectively. (b) LSV curves, (c) corresponding overpotential values, (d) Tafel plots, (e) TOF values, (f) Nyquist plots, (g) long-term stability assessment, and (h) comparison of overpotential with previously reported catalysts.

Further examination of the electrochemical surface area (ECSA) and double-layer capacitance (Cdl) provided critical insights into the active surface properties of the HE-Cr(NiFeCoV)2O4 catalyst.45 CV measurements at varying scan rates (Fig. S11a–d) were used to determine the ECSA and Cdl of the catalysts. As shown in Fig. S12 and S13, HE-Cr(NiFeCoV)2O4 exhibited a higher Cdl of 18.30 mF cm−2 and ECSA of 1.83 cm2 compared to CrCo2O4 (16.62 mF cm−2 and 1.66 cm2) and commercial IrO2 (14.02 mF cm−2 and 1.40 cm2). These results indicate that HE-Cr(NiFeCoV)2O4 possesses a greater number of accessible active sites, attributed to its high-entropy structure, which enhances both the intrinsic activity and availability of catalytically active regions during OER.30,46 In addition to catalytic activity, long-term stability of electrocatalysts is a critical parameter for real-world applications. Long-term stability testing results of the HE-Cr(NiFeCoV)2O4 catalyst are displayed in Fig. 3g. Notably, during the prolonged stability test conducted at 50 mA cm−2 for over 100 h, HE-Cr(NiFeCoV)2O4 maintained a nearly stable potential throughout continuous OER operation, highlighting its excellent long-term stability. Post-stability analyses, including SEM, EDS, and XPS, further confirmed the robustness of the HE-Cr(NiFeCoV)2O4 electrode before and after OER stability testing at 50 mA cm−2 over 100 h. SEM and EDS images (Fig. S14) showed that the surface morphology and elemental composition of the HE-Cr(NiFeCoV)2O4 electrode were retained after extended OER operation. Additionally, XPS plots (Fig. S15a–f) indicated slight increases in Cr4+ intensities and shifts in the Cr 2p, Ni 2p, Fe 2p, Co 2p, V 2p, and O 1s peaks of the HE-Cr(NiFeCoV)2O4 electrode following OER stability testing, suggesting electrochemical modifications to its surface during continuous OER. This surface modification likely contributed to the enhanced OER performance of the HE-Cr(NiFeCoV)2O4 catalyst. Remarkably, Fig. 3h and Table S1 illustrate that the HE-Cr(NiFeCoV)2O4 catalyst exhibited the lowest overpotential among many recently reported OER catalysts. Overall, these comprehensive electrochemical results highlight the exceptional OER performance of the HE-Cr(NiFeCoV)2O4 catalyst, which combines low overpotential with excellent stability. These advantages can be primarily attributed to the synergistic effects of its multicomponent composition, optimized electronic structure, large surface area, enhanced electrical conductivity, increased configurational entropy, and improved surface-active sites. Thus, the CO2 laser-engineered HE-Cr(NiFeCoV)2O4 catalyst is a promising candidate for OER in water electrolysis, advancing sustainable H2 production.

3.3. Electrochemical performance of HE-Cr(NiFeCoV)2O4 toward OWS

Based on the exceptional OER performance of the HE-Cr(NiFeCoV)2O4 catalyst, an alkaline OWS electrolyzer was assembled using HE-Cr(NiFeCoV)2O4 as the anode and standard Pt/C as the cathode in 1 M KOH electrolyte. A schematic illustration of the assembled OWS electrolyzer with the HE-Cr(NiFeCoV)2O4‖Pt/C system is provided in Fig. 4a. As shown in Fig. 4b, the LSV profiles recorded at 5 mV s−1 revealed that the HE-Cr(NiFeCoV)2O4‖Pt/C system exhibited superior OWS performance compared to the benchmark IrO2‖Pt/C system. At current densities of 10, 50, and 100 mA cm−2, the HE-Cr(NiFeCoV)2O4‖Pt/C system required substantially lower cell voltages of 1.57, 1.77, and 1.96 V, respectively, compared to the 1.61, 1.87, and 2.15 V for the IrO2‖Pt/C system under identical conditions (Fig. 4c). This improved OWS efficiency of the HE-Cr(NiFeCoV)2O4‖Pt/C system can be primarily attributed to the improved OER performance facilitated by the HE-Cr(NiFeCoV)2O4 anode. Furthermore, gas chromatography was employed to quantitatively evaluate H2 production by the HE-Cr(NiFeCoV)2O4‖Pt/C system. As illustrated in Fig. 4d, the HE-Cr(NiFeCoV)2O4‖Pt/C system produced 0.592 mmol of H2 within 60 min, substantially exceeding the 0.516 mmol generated by the IrO2‖Pt/C system under identical conditions. These results confirm the superior catalytic efficiency of the HE-Cr(NiFeCoV)2O4 catalyst in facilitating OWS for efficient H2 production. Long-term stability testing was then conducted by maintaining the HE-Cr(NiFeCoV)2O4‖Pt/C system at 50 mA cm−2 for 100 h. The results, presented in Fig. 4e, exhibited minimal performance degradation, with only a 100 mV increase in cell voltage over the continuous 100 h OWS operation. This prolonged operational stability demonstrates the robustness of the HE-Cr(NiFeCoV)2O4 anode under alkaline conditions, making it suitable for practical water electrolysis applications. Finally, a comparative analysis with previously reported OWS systems (Fig. 4f and Table S2), highlights the superior performance of the HE-Cr(NiFeCoV)2O4‖Pt/C system, particularly in achieving low cell voltages. Therefore, the combination of high catalytic activity, low energy consumption, and long-term stability positions the HE-Cr(NiFeCoV)2O4-based system as a promising candidate for sustainable and efficient H2 production through OWS.
image file: d5ta01573a-f4.tif
Fig. 4 OWS performance of Cr(NiFeCoV)2O4‖Pt/C and IrO2‖Pt/C electrolyzers: (a) schematic of the OWS system, (b) LSV curves, (c) corresponding cell voltages, (d) quantitative analysis of H2 production at 2 V, (e) long-term stability evaluation, and (f) comparison of cell voltage requirements with previously reported OER catalysts for OWS systems.

3.4. In situ Raman and theoretical probes on the HE-Cr(NiFeCoV)2O4 catalyst

Advanced in situ Raman spectroscopy was employed to monitor the real-time formation of surface-active intermediates on the HE-Cr(NiFeCoV)2O4 catalyst during OER. As schematically depicted in Fig. 5a, a 532 nm laser was focused on the surface of the HE-Cr(NiFeCoV)2O4 electrode during the OER process to capture the real-time Raman spectra at potentials ranging from 1.2 to 1.7 V vs. RHE. The recorded spectra (Fig. 5b) revealed two distinct Raman peaks at 699 and 802 cm−1, corresponding to FeOOH. The intensities of these peaks increased as the potential rose from 1.2 to 1.7 V vs. RHE, verifying the formation of surface-active FeOOH on the HE-Cr(NiFeCoV)2O4 electrode during the OER process.47,48 This surface transformation is pivotal to enhancing the overall catalytic OER performance of the HE-Cr(NiFeCoV)2O4 catalyst.
image file: d5ta01573a-f5.tif
Fig. 5 (a) Schematic representation of the in situ Raman spectroscopy setup. (b) In situ Raman spectra of the HE-Cr(NiFeCoV)2O4 catalyst during OER. (c) DFT-derived top-view crystal structures of the CrCo2O4 catalyst. (d) DFT-derived top-view crystal structures of the HE-Cr(NiFeCoV)2O4 catalyst. (e) Charge density difference map of the HE-Cr(NiFeCoV)2O4 catalyst. (f) PDOS plots for the CrCo2O4 catalyst. (g) PDOS plots for the HE-Cr(NiFeCoV)2O4 catalyst. (h) TDOS comparison between the CrCo2O4 and HE-Cr(NiFeCoV)2O4 catalysts.

To gain a deeper understanding of how multiple metal components influence the electronic structure and OER performance of the HE-Cr(NiFeCoV)2O4 catalyst, DFT calculations were performed. The optimized top- and side-view crystal structures of CrCo2O4 (Fig. 5c and S16) and HE-Cr(NiFeCoV)2O4 (Fig. 5d and S17) confirmed the successful incorporation of Cr, Ni, Fe, Co, and V into a single spinel oxide structure. The charge density difference plots (Fig. 5e and S18) revealed considerable redistribution of charge around the metal centers, indicating electronic modulation throughout the HE-Cr(NiFeCoV)2O4 structure. This is ascribed to the incorporation of multiple transition metal cations, which further enhances the electronic properties of the catalyst and is consistent with its superior OER performance.44 Moreover, the partial and total density of states (PDOS and TDOS) plots (Fig. 5f–h) reveal a substantially higher TDOS near the Fermi level for HE-Cr(NiFeCoV)2O4 compared to CrCo2O4. This enhancement primarily arises from the combined contributions of Cr(d), Ni(d), Fe(d), Co(d), and V(d) orbitals, suggesting improved electrical conductivity that facilitates rapid charge transport during OER. These theoretical results further corroborate the experimentally observed enhancement in OER performance of HE-Cr(NiFeCoV)2O4.49

To evaluate the intrinsic OER activity of various metal sites, we calculated the Gibbs free energy changes associated with the adsorption and transformation of key OER intermediates (*OH, O*, and *OOH) on the Cr, Ni, Fe, Co, and V sites of the HE-Cr(NiFeCoV)2O4 surface. The 4e OER pathway at these active sites, along with the corresponding Gibbs free energy diagrams, is illustrated in Fig. 6a–e. The reaction mechanism proceeds via the adsorption and dissociation of H2O to form the *OH intermediates, followed by the sequential formation of *O and *OOH species. The final step involves the evolution of O2, thereby regenerating the active site. Gibbs free energy diagrams calculated at U = 0 V and U = 1.23 V indicate that the formation of *OOH constitutes the potential-determining step (PDS) across all active sites. Among these, the Fe site exhibits the lowest overpotential of 0.45 V, suggesting it is the most favorable active site for OER. This theoretical finding is consistent with Raman spectroscopy results, which demonstrate the in situ formation of FeOOH during OER, likely due to progressive oxidation at the Fe sites. To further investigate this phenomenon, we constructed a model featuring a surface FeOOH layer on HE-Cr(NiFeCoV)2O4 (Fig. S19), optimized the adsorption of OER intermediates on the Fe site of this modified structure, and calculated the corresponding Gibbs free energy profile (Fig. 6f). Notably, the presence of FeOOH further decreased the OER overpotential from 0.45 V to 0.41 V, emphasizing the critical role of surface-formed FeOOH in enhancing OER activity. Collectively, these DFT results highlight the synergistic effects of multi-metallic sites in modulating the electronic structure and reducing reaction energy barriers, thereby conferring superior OER performance to the HE-Cr(NiFeCoV)2O4 catalyst.1,50 Consequently, the continuous-wave CO2 laser-synthesized HE-Cr(NiFeCoV)2O4 catalyst emerges as a promising candidate for water electrolysis, facilitating efficient H2 production.


image file: d5ta01573a-f6.tif
Fig. 6 (a–e) OER mechanism and corresponding Gibbs free energy diagrams for the (a) Cr site, (b) Ni site, (c) Fe site, (d) Co site, and (e) V site of the HE-Cr(NiFeCoV)2O4 catalyst. (f) OER mechanism along with the corresponding Gibbs free energy diagrams at the Fe-site of the surface-formed FeOOH on the HE-Cr(NiFeCoV)2O4 catalyst.

4. Conclusions

Herein, we present a novel, simple, and energy-efficient continuous-wave CO2 laser technology for synthesizing spinel-structured HE-Cr(NiFeCoV)2O4 HEOs via a rapid thermal-shock approach. This approach facilitates the efficient transformation of metal hydroxide precursors into HE-Cr(NiFeCoV)2O4 HEOs with a homogeneously distributed multicomponent structure within a short processing time. This method offers several advantages over traditional synthesis techniques, including substantial time and energy savings, prevention of impurity formation, and reduced production costs. Notably, the HE-Cr(NiFeCoV)2O4 catalyst demonstrated superior OER performance, requiring a lower overpotential of 284 mV at 10 mA cm−2 in 1 M KOH electrolyte, outperforming the bimetallic CrCo2O4 catalyst (404 mV at 10 mA cm−2). The catalyst also exhibited excellent long-term stability, maintaining consistent performance over 100 h of continuous OER operation at 50 mA cm−2. When integrated HE-Cr(NiFeCoV)2O4 as the anode in an OWS electrolyzer with Pt/C as the cathode, the system achieved efficient H2 production with an exceptionally low cell voltage of 1.57 V at 10 mA cm−2 and sustained remarkable stability over 100 h at 50 mA cm−2. These findings underscore the potential of HE-Cr(NiFeCoV)2O4 catalyst as a highly suitable candidate for OER in large-scale H2 production. Further insights into its superior OER performance were provided via in situ Raman spectroscopy and theoretical DFT analyses. Raman studies confirmed the formation of active FeOOH surface species, notably contributing to the enhanced OER performance. DFT calculations revealed that the multi-element composition of HE-Cr(NiFeCoV)2O4 optimized its electronic structure, lowering energy barriers for reaction intermediates during OER. The synergistic interaction between Cr, Ni, Fe, Co, and V cations within a single oxide lattice endows the catalyst with unique physicochemical properties that boost OER performance. Overall, this study not only demonstrates the potential of HEOs as robust and efficient electrocatalysts for sustainable energy applications but also paves the way for designing tailored HEOs for multifaceted catalytic systems using the facile and innovative continuous-wave CO2 laser technology.

Data availability

The data supporting this article have been included as part of the ESI.

Conflicts of interest

No conflict to declare.

Acknowledgements

This research was supported by the Korea Basic Science Institute (National research Facilities and Equipment Center) grant funded by the Ministry of Education (No. 2019R1A6C1010042 and RS-2024-00434932) and the Glocal University 30 Project Fund of Gyeongsang National University in 2024. The authors acknowledge financial support from the National Research Foundation of Korea (NRF) (2022R1A2C2010686 and 2022R1I1A1A01073299). MU acknowledges the funding support from the Researchers Supporting Project (No. RSPD2025R682), King Saud University, Riyadh, Saudi Arabia.

References

  1. D. Wang, Y. Chu, Y. Wu, M. Zhu, L. Pan, R. Li, Y. Chen, W. Wang, N. Mitsuzaki and Z. Chen, J. Mater. Chem. A, 2024, 12, 33680–33688 RSC.
  2. Y. Li, A. Feng, L. Dai, B. Xi, X. An, S. Xiong and C. An, Adv. Funct. Mater., 2024, 34, 2316296 CrossRef CAS.
  3. Z. Zhao, J. Sun, X. Li, S. Qin, C. Li, Z. Zhang, Z. Li and X. Meng, Nat. Commun., 2024, 15, 7475 CrossRef CAS PubMed.
  4. V. Maheskumar, A. Min, A. Kumar, R. A. Senthil, C. J. Moon and M. Y. Choi, Small, 2024, 20, 2403314 CrossRef CAS PubMed.
  5. Z. Li, G. Lin, L. Wang, H. Lee, J. Du, T. Tang, G. Ding, R. Ren, W. Li, X. Cao, S. Ding, W. Ye, W. Yang and L. Sun, Nat. Catal., 2024, 7, 944–952 CrossRef CAS.
  6. M. Hou, L. Zheng, D. Zhao, X. Tan, W. Feng, J. Fu, T. Wei, M. Cao, J. Zhang and C. Chen, Nat. Commun., 2024, 15, 1342 CrossRef CAS.
  7. S. Zuo, Z.-P. Wu, D. Xu, R. Ahmad, L. Zheng, J. Zhang, L. Zhao, W. Huang, H. Al Qahtani, Y. Han, L. Cavallo and H. Zhang, Nat. Commun., 2024, 15, 9514 CrossRef CAS.
  8. S. Qian, T. Jiang, J. Wang, W. Yuan, D. Jia, N. Cheng, H. Xue, Z. Xu, R. Gautier and J. Tian, ACS Catal., 2024, 14, 18690–18700 CrossRef CAS.
  9. S. Jung, R. A. Senthil, A. Min, A. Kumar, C. J. Moon, G. H. Jeong, T. W. Kim and M. Y. Choi, J. Mater. Chem. A, 2024, 12, 8694–8706 RSC.
  10. D. H. Lee, R. Kerkar, D. Arumugam, J. Theerthagiri, S. Ramasamy, S. Kheawhom and M. Y. Choi, J. Mater. Chem. A, 2024, 12, 30269–30278 RSC.
  11. Z.-D. He, R. Tesch, M. J. Eslamibidgoli, M. H. Eikerling and P. M. Kowalski, Nat. Commun., 2023, 14, 3498 CrossRef CAS PubMed.
  12. P. Zhai, C. Wang, Y. Zhao, Y. Zhang, J. Gao, L. Sun and J. Hou, Nat. Commun., 2023, 14, 1873 CrossRef CAS.
  13. A. P. Sundar Rajan, R. A. Senthil, C. J. Moon, A. Kumar, A. Min, M. Ubaidullah and M. Y. Choi, Chem. Eng. J., 2024, 502, 157848 CrossRef CAS.
  14. S. Jung, R. A. Senthil, A. Min, A. Kumar, C. J. Moon and M. Y. Choi, Small Methods, 2024, 8, 2301628 CrossRef CAS PubMed.
  15. Q. Deng, H. Li, K. Pei, L. W. Wong, X. Zheng, C. S. Tsang, H. Chen, W. Shen, T. H. Ly, J. Zhao and Q. Fu, ACS Nano, 2024, 18, 33718–33728 CrossRef CAS PubMed.
  16. R. A. Senthil, S. Jung, A. Min, A. Kumar, C. J. Moon, M. Singh and M. Y. Choi, ACS Catal., 2024, 14, 3320–3335 CrossRef CAS.
  17. R. A. Senthil, S. Jung, A. Min, C. J. Moon and M. Y. Choi, Chem. Eng. J., 2023, 475, 146441 CrossRef CAS.
  18. J. Baek, M. D. Hossain, P. Mukherjee, J. Lee, K. T. Winther, J. Leem, Y. Jiang, W. C. Chueh, M. Bajdich and X. Zheng, Nat. Commun., 2023, 14, 5936 CrossRef CAS PubMed.
  19. A. Sarkar, L. Velasco, D. Wang, Q. Wang, G. Talasila, L. de Biasi, C. Kübel, T. Brezesinski, S. S. Bhattacharya, H. Hahn and B. Breitung, Nat. Commun., 2018, 9, 3400 CrossRef.
  20. C. Triolo, K. Moulaee, A. Ponti, G. Pagot, V. Di Noto, N. Pinna, G. Neri and S. Santangelo, Adv. Funct. Mater., 2024, 34, 2306375 CrossRef CAS.
  21. K. Wang, W. Hua, X. Huang, D. Stenzel, J. Wang, Z. Ding, Y. Cui, Q. Wang, H. Ehrenberg, B. Breitung, C. Kübel and X. Mu, Nat. Commun., 2023, 14, 1487 CrossRef CAS PubMed.
  22. S. C. Karthikeyan, S. Ramakrishnan, S. Prabhakaran, M. R. Subramaniam, M. Mamlouk, D. H. Kim and D. J. Yoo, Small, 2024, 20, 2402241 CrossRef CAS.
  23. C. E. Park, R. A. Senthil, G. H. Jeong and M. Y. Choi, Small, 2023, 19, 2207820 CrossRef CAS.
  24. X. Wang, R. Lu, S. Gong, S. Yang, W. Wang, Z. Sun, X. Zhang, J. Liu and X. Lv, Chem. Eng. J., 2025, 503, 158488 CrossRef CAS.
  25. Y. Liu, C. Ye, L. Chen, J. Fan, C. Liu, L. Xue, J. Sun, W. Zhang, X. Wang, P. Xiong and J. Zhu, Adv. Funct. Mater., 2024, 34, 2314820 CrossRef CAS.
  26. S. S. Aamlid, M. Oudah, J. Rottler and A. M. Hallas, J. Am. Chem. Soc., 2023, 145, 5991–6006 CrossRef CAS.
  27. Y. Huang, D. Wang, Y. Yu, Z. Li, X. Wen and Z. Wang, Energy Fuels, 2024, 38, 13661–13684 CrossRef CAS.
  28. J. Wang, S. Sun, S. Xi, Y. Sun, S. J. H. Ong, Z. W. Seh and Z. J. Xu, J. Phys. Chem. C, 2024, 128, 4978–4987 CrossRef CAS.
  29. M. Zhang, J. Ye, Y. Gao, X. Duan, J. Zhao, S. Zhang, X. Lu, K. Luo, Q. Wang, Q. Niu, P. Zhang and S. Dai, ACS Nano, 2024, 18, 1449–1463 Search PubMed.
  30. K. Miao, W. Jiang, Z. Chen, Y. Luo, D. Xiang, C. Wang and X. Kang, Adv. Mater., 2024, 36, 2308490 CrossRef CAS PubMed.
  31. Y. Lee, J. Theerthagiri, N. Yodsin, A. Min, C. J. Moon, S. Jungsuttiwong and M. Y. Choi, Angew. Chem., Int. Ed., 2024, 63, e202413774 CrossRef CAS.
  32. D. Van Lam, M. Sohail, J.-H. Kim, H. J. Lee, S. O. Han, J. Shin, D. Kim, H. Kim and S.-M. Lee, ACS Appl. Mater. Interfaces, 2020, 12, 39154–39162 CrossRef CAS.
  33. K. Kashihara, Y. Uto and T. Nakajima, Sci. Rep., 2018, 8, 14719 CrossRef PubMed.
  34. T. X. Nguyen, J. Patra, J.-K. Chang and J.-M. Ting, J. Mater. Chem. A, 2020, 8, 18963–18973 RSC.
  35. D.-R. Kong, Y. Gao, Y.-Y. Xin, B. Li, X.-F. Zhang, Z.-P. Deng, L.-H. Huo and S. Gao, Ceram. Int., 2023, 49, 19206–19215 CrossRef CAS.
  36. S. Jung, R. A. Senthil, C. J. Moon, N. Tarasenka, A. Min, S. J. Lee, N. Tarasenko and M. Y. Choi, Chem. Eng. J., 2023, 468, 143717 CrossRef CAS.
  37. Y. Zhu, Q. Xiang, L. Guo, S. Lu, F. Duan, M. Du and H. Zhu, Colloids Surf., A, 2024, 686, 133315 CrossRef CAS.
  38. T. X. Nguyen, Y.-C. Liao, C.-C. Lin, Y.-H. Su and J.-M. Ting, Adv. Funct. Mater., 2021, 31, 2101632 CrossRef CAS.
  39. S. E. Michaud, M. T. Riehs, W.-J. Feng, C.-C. Lin and C. C. L. McCrory, Chem. Commun., 2021, 57, 883–886 RSC.
  40. R. Muruganantham, J.-S. Lu and W.-R. Liu, Polymers, 2020, 12, 555 CrossRef CAS.
  41. J. Mu, Z. Zhao, X.-W. Gao, Z.-M. Liu, W.-B. Luo, Z. Sun, Q.-F. Gu and F. Li, Adv. Energy Mater., 2024, 14, 2303558 CrossRef CAS.
  42. Z. Chen, W. Gong, J. Wang, S. Hou, G. Yang, C. Zhu, X. Fan, Y. Li, R. Gao and Y. Cui, Nat. Commun., 2023, 14, 5363 CrossRef CAS PubMed.
  43. M. I. Abdullah, Y. Fang, X. Wu, M. Hu, J. Shao, Y. Tao and H. Wang, Nat. Commun., 2024, 15, 10587 CrossRef CAS.
  44. R. Liu, Y. Yan, L. Dun, T. Yang, B. Qin, P. Wang, W. Cai, S. Liu and X. Zheng, Mat. Tod. Catal., 2025, 8, 100086 Search PubMed.
  45. V. Maheskumar, A. Min, C. J. Moon, R. A. Senthil and M. Y. Choi, Small Struct., 2023, 4, 2300212 CrossRef CAS.
  46. T. Li, Y. Yao, B. H. Ko, Z. Huang, Q. Dong, J. Gao, W. Chen, J. Li, S. Li, X. Wang, R. Shahbazian-Yassar, F. Jiao and L. Hu, Adv. Funct. Mater., 2021, 31, 2010561 CrossRef CAS.
  47. Z. Liu, N. Zhang and Y. Xiong, J. Phys. Chem. C, 2024, 128, 13651–13665 CrossRef CAS.
  48. X. Xu, Y. Zhang, Y. Chen, C. Liu, W. Wang, J. Wang, H. Huang, J. Feng, Z. Li and Z. Zou, Proc. Natl. Acad. Sci. U. S. A., 2022, 119, e2205562119 CrossRef CAS PubMed.
  49. X. Zhang, X. Wang and X. Lv, ChemSusChem, 2025, 18, e202401663 CrossRef CAS.
  50. K. L. Svane and J. Rossmeisl, Angew. Chem., Int. Ed., 2022, 61, e202201146 CrossRef CAS.

Footnotes

Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d5ta01573a
These authors 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.