Next Article in Journal
Surface Hardness of Polished Dental Zirconia: Influence of Polishing and Yttria Content on Morphology, Phase Composition, and Microhardness
Previous Article in Journal
Study on Mechanical Properties of Composite Basalt Fiber 3D-Printed Concrete Based on 3D Meso-Structure
Previous Article in Special Issue
SiOx-Based Anode Materials with High Si Content Achieved Through Uniform Nano-Si Dispersion for Li-Ion Batteries
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Multiphase NiCoFe-Based LDH for Electrocatalytic Sulfion Oxidation Reaction Assisting Efficient Hydrogen Production

by
Zengren Liang
1,
Yong Nian
1,
Hao Du
2,
Peng Li
2,
Mei Wang
1,* and
Guanshui Ma
3,4,*
1
School of Materials Science and Engineering, North University of China, Taiyuan 030051, China
2
School of Energy and Power Engineering, North University of China, Taiyuan 030051, China
3
State Key Laboratory of Advanced Marine Materials, Ningbo Institute of Materials Technology and Engineering, Chinese Academy of Sciences, Ningbo 315201, China
4
Zhejiang Key Laboratory of Extreme Environmental Material Surfaces and Interfaces, Ningbo Institute of Materials Technology and Engineering, Chinese Academy of Sciences, Ningbo 315201, China
*
Authors to whom correspondence should be addressed.
Materials 2025, 18(14), 3377; https://doi.org/10.3390/ma18143377
Submission received: 20 June 2025 / Revised: 13 July 2025 / Accepted: 16 July 2025 / Published: 18 July 2025
(This article belongs to the Special Issue High-Performance Materials for Energy Conversion)

Abstract

Sulfion oxidation reaction (SOR) has great potential in replacing oxygen evolution reaction (OER) and boosting highly efficient hydrogen evolution. The development of highly active and stable SOR electrocatalysts is crucial for assisting hydrogen production with low energy consumption. In this work, multiphase NiCoFe-based layered double hydroxide (namely NiCoFe-LDH) has been synthesized via a facile seed-assisted heterogeneous nucleation method. Benefiting from its unique microsized hydrangea-like structure and synergistic active phases, the catalyst delivers substantial catalytic interfaces and reactive centers for SOR. Consequently, NiCoFe-LDH electrode achieves a remarkably low potential of 0.381 V at 10 mA cm−2 in 1 M KOH + 0.1 M Na2S, representing a significant reduction of 0.98 V compared to conventional OER. Notably, under harsh industrial conditions (6 M KOH + 0.1 M Na2S, 80 °C), the electrolysis system based on NiCoFe-LDH||NF pair exhibits a cell potential of only 0.71 V at 100 mA cm−2, which shows a greater decreasing amplitude of 1.05 V compared with that of traditional OER/HER systems. Meanwhile, the NiCoFe-LDH||NF couple could maintain operational stability for 100 h without obvious potential fluctuation, as well as possessing a lower energy consumption of 1.42 kWh m−3 H2. Multiphase eletrocatalysis for SOR could indeed produce hydrogen with low-energy consumption.

Graphical Abstract

1. Introduction

Utilizing renewable energy sources (such as electricity converted from solar and wind power) to produce hydrogen via water electrolysis [1,2,3], as a substitute for traditional fossil fuels [4,5,6], represents an effective pathway to address both fossil fuel depletion and global climate change [7,8]. However, in traditional water electrolysis, the anodic oxygen evolution reaction (OER: 4OH → O2 + 2H2O + 4e, 1.23 V vs. RHE), involving the breaking of O-H bonds and the formation of O=O bonds, requires the transfer of four electrons, which renders the reaction thermodynamically sluggish and results in elevated cell voltage and excessive energy consumption [9,10]. Currently, an extremely promising and efficient hydrogen production strategy involves introducing additional electrolytes into the electrolyte solution. The anodic oxidation reaction using these additional electrolytes is utilized to replace the slow OER. The oxidation potentials of these additional electrolytes are much lower than that of the traditional OER, for example, urea oxidation reaction (UOR, 0.37 V vs. RHE) [11,12], glucose oxidation reaction (GOR, 0.05 V vs. RHE) [13,14], methanol oxidation reaction (MOR, 0.016 V vs. RHE) [15,16], hydrazine oxidation reaction (HzOR, −0.33 V vs. RHE), [17,18] etc., which can significantly reduce the overpotential of the anode reaction, thereby effectively lowering the overall electrolysis voltage of the electrolytic system. This means that under the same electrolysis voltage, a higher current density and hydrogen production rate can be achieved. It is particularly noteworthy that the sulfion oxidation reaction (SOR, S2− → S + 2e, −0.48 V vs. RHE) involves only two electron transfers during the reaction process, refs. [9,19,20] resulting in a faster reaction kinetics and a lower oxidation potential compared to the aforementioned organic compounds (Table S1). This means that SOR has a more significant advantage in reducing the overpotential of the anode reaction. Therefore, if SOR is coupled with the cathode’s HER, even when operating with a significantly reduced total voltage, the electrolytic system can still maintain efficient HER at the cathode, thereby achieving the goal of reducing energy consumption in hydrogen production.
Consequently, designing a high-performance SOR catalyst is critical. It is noteworthy that transition metal (e.g., Ni, Co, Fe)-based layered double hydroxides (LDHs) demonstrate significant potential in the field of electrocatalytic SOR, owing to their simple synthesis process, abundant active sites, excellent electron transfer properties, and high specific surface area [21,22,23]. For example, the NiFe-LDH/FeNi2S4/IF constructed by Ai’s team exhibits high mass-transfer efficiency in SOR, benefiting from the superhydrophilic surface and high specific surface area provided by the NiFe-LDH phase [24]. Jiang’s research group reported NiFe-LDH/Cu-IF, where the unique structure of the NiFe-LDH phase—characterized by interconnected porous ultrathin nanosheets with a large specific surface area—significantly contributes to the enhancement of electrocatalytic SOR activity [25].
Inspired by the above, through facile seed-assisted heterogeneous nucleation method [26], this work synthesizes a NiCoFe-LDH catalyst, featuring a unique hydrangea-like microstructure and multiple synergistic active phases (NiCoFe-based (oxy)hydroxides) that endow it with an exceptionally large specific surface area rich in catalytic interfaces and reaction centers, thus delivering exceptional catalytic performance for the SOR. In 1 M KOH and 0.1 M Na2S, the NiCoFe-LDH catalyst required an electrolysis potential of only 0.381 V (vs. RHE) at 10 mA cm−2, representing a reduction of 0.98 V compared to traditional OER under identical conditions. Long-term stability testing reveals that NiCoFe-LDH catalyst maintains stable operation for 100 h with minimal potential fluctuation. Under industrial operating conditions (6 M KOH + 0.1 M Na2S, 80 °C), an SOR/HER electrolysis system based on the NiCoFe-LDH||NF pair requires only 0.71 V to sustain a current density of 100 mA cm−2, representing a reduction of 1.05 V compared to conventional OER/HER system. Additionally, the energy consumption for H2 production in this system has been reduced to 1.42 kWh m−3 H2, which is far below the theoretical minimum energy consumption of OER/HER system (2.94 kWh m−3 H2) [9]. From a theoretical perspective, the adsorptions of S onto Fe2O3, CoOOH and Ni(OH)2 are all spontaneous, and the corresponding intensities are as follows: CoOOH > Ni(OH)2 > Fe2O3. All the above signifies that the SOR/HER electrolysis system based on NiCoFe-LDH||NF pair can achieve the goal of the energy-efficient production of green hydrogen.

2. Experimental Section

2.1. Materials Preparation

The NiCoFe-LDH catalyst was grown in situ on nickel foam (NF, ≥99.99%, thickness of 2  mm, Sinero, Suzhou, China) via the seed-assisted heterogeneous nucleation method (Scheme 1) [26]. Firstly, the nickel foam (5 × 5 cm2) was ultrasonically cleaned in acetone and anhydrous ethanol to remove surface oils and other impurities. Secondly, it was immersed in 3 M dilute hydrochloric acid for 20 min to remove the surface oxide layer, then rinsed with deionized water and dried for later use. Next, 14.4 g of nickel (II) nitrate hexahydrate (Ni(NO3)2·6H2O, ≥99.0%, Aladdin Industrial, Shanghai, China) and 3.6 g of cobalt (II) nitrate hexahydrate (Co(NO3)2·6H2O, ≥99.0%, Aladdin Industrial, Shanghai, China) were dissolved in 240 mL of isopropanol to form a mixed solution. Separately, 1.95 g of iron (II) sulfate heptahydrate (FeSO4·7H2O, ≥99.0%, Aladdin Industrial, Shanghai, China) was dissolved in 80 mL of deionized water. Subsequently, these two solutions were combined and subjected to continuous vigorous stirring for approximately 10 min, followed by natural sedimentation for 5 min to form a stable suspension of NiCoFe nanoparticle seeds. Previous reports indicate that nanoparticles in the suspension facilitate the subsequent in situ growth of the catalyst [26]. Following this, the pretreated NF was immersed in the NiCoFe nanoparticle seed suspension at 25 °C for 48 h. Afterwards, it was rinsed with deionized water and dried in a vacuum oven at 25 °C for 12 h, resulting in the final NiCoFe-LDH catalyst.

2.2. Physicochemical Characterization

The phase composition and crystal structure information of the NiCoFe-LDH catalyst powder were characterized using X-ray diffraction (XRD, DX-2700BH, Haoyuan Instrument Co., Ltd., Dandong, China) with Cu Kα radiation (λ = 0.154 nm) operating at 40 kV and 30 mA. The concentration of vacancy defects was evaluated by electron paramagnetic resonance (EPR, JES-FA200, JEOL, Tokyo, Japan). The microstructure of the material was imaged using scanning electron microscopy (SEM, JSM-7900F, JEOL, Tokyo, Japan).

2.3. Electrochemical Measurements

The as-prepared NiCoFe-LDH is a self-supporting electrode that can be directly used for testing and does not need a binder, while the control RuO2 (99.9%, Macklin reagent, Shanghai, China) electrode should be prepared as follows: 4 mg RuO2 powder and 1 mg carbon black (Vulcan XC-72, Cabot, Boston, MA, USA) were dispersed in 300 μL isopropanol, followed by adding 100 μL Nafion solution (5 wt%, DuPont, Nogales, AZ, USA). The mixture was ultrasonicated to form a homogeneous RuO2 catalyst ink. Subsequently, 50 μL of the RuO2 catalyst ink was loaded onto a 1 cm2 nickel foam (NF) substrate, resulting in a mass loading of approximately 0.5 mg cm−2.
For the half-reaction electrochemical tests, measurements were performed at room temperature using a three-electrode system connected to an electrochemical workstation (CS2350H, CorrTest, Wuhan, China). The catalyst electrode served as the working electrode (WE), a platinum sheet as the counter electrode (CE), and an Hg/HgO electrode as the reference electrode (RE). All half-reaction potentials were converted to the reversible hydrogen electrode (RHE) scale according to the following equation:
ERHE = EHg/HgO + 0.098 + 0.059 × pH
The potentiodynamic scan test, one of the half-reaction tests, was performed at a scanning rate of 0.05 mV/s to evaluate the corrosive effect of electrolytes containing different concentrations of Sodium sulfide hexahydrate (Na2S·6H2O, ≥98.0%, Aladdin Industrial, Shanghai, China) on the NiCoFe-LDH electrode. In addition, half-reaction tests also included measurements of alkaline OER and SOR in 1.0 M KOH (≥85.0%, Damao Chemical Reagent, Tianjin, China) with and without Na2S present (i.e., in 1.0 M KOH and 1.0 M KOH + 0.1 M Na2S). OER and SOR activities were evaluated by linear-sweep voltammetry (LSV) at a scan rate of 5 mV s−1. All test results were recorded without iR compensation. Tafel slopes were extracted from the LSV data according to the following equation:
η = b × log|j| + a
where η, b, j, and a represent the overpotential, Tafel slope, current density, and Tafel constant, respectively. Electrochemical impedance spectroscopy (EIS) measurements were conducted in the frequency range from 105 Hz to 0.01 Hz at a constant potential of 1.4 V vs. RHE (in 1.0 M KOH). SOR stability was evaluated through chronopotentiometry (CP) at a constant current density of 10 mA cm−2. The ECSA reflects intrinsic catalyst properties and can be calculated using the double-layer capacitance (Cdl) values of catalyst according to the following equation:
ECSA = Cdl/Cs = Cdl/(40 μ−2per2ECSA)
Among these, the Cdl value was determined by conducting cyclic voltammetry (CV) measurements within the non-Faradaic potential region (OCP ± 0.05 V vs. RHE), at various scan rates ranging from 5 to 100 mV s−1. The slope was then extracted from the plot of |ja − jc|/2 versus scan rate. Cs is typically chosen as 40 µF cm−2. Furthermore, SOR-coupled HER water electrolysis was tested in an anion exchange membrane (AEM) electrolyzer. The electrolyzer employed NiCoFe-LDH as the anode and nickel foam (NF) as the cathode, both with an electrode area of 2 cm × 2 cm. The electrolyte was 1.0 M KOH + 0.1 M Na2S at 80 °C. Polarization curves were recorded at a scan rate of 5 mV s−1, and stability was measured at a current density of 100 mA cm−2.

2.4. Theoretical Calculation Method

Theoretical calculations were performed using density functional theory (DFT) as implemented in the Vienna ab initio simulation package (VASP 6.3.0). The Perdew–Burke–Ernzerhof (PBE) functional was employed. To address the challenges of highly localized and strongly correlated electronic systems, a DFT + U correction was applied. The Hubbard U values were set to 3.4 eV for Ni, 3.4 eV for Co, and 3.3 eV for Fe. To prevent interactions between periodic images of the slab, a vacuum layer thickness of 15 Å was introduced, with 7.5 Å above and 7.5 Å below the catalyst slab. Additional computational parameters included a plane-wave cutoff energy of 400 eV, a self-consistent field (SCF) convergence threshold of 10−5 eV, an energy convergence criterion of 0.03 eV Å−1 for ionic relaxation, and a k-point mesh of 5 × 5 × 1.

3. Results and Discussion

3.1. Microstructure Analysis

To avoid the characteristic peaks of the NiCoFe-LDH catalyst being masked by those of the nickel foam (NF) substrate, the NiCoFe-LDH powder was scraped off the NF surface. Phase analysis of the catalyst powder was then performed using X-ray diffraction (XRD). The phase analysis revealed that NiCoFe-LDH was not a single phase, but rather composed of Ni(OH)2 (PDF#97-016-9978), CoO(OH) (PDF#99-000-1533), Fe2O3 (PDF#97-003-6281) and Ni (PDF#04-001-1136) phases (Figure 1a). Thereinto, the detection of Ni was due to the inadvertent scraping of the surface layer of the NF substrate during the removal of the NiCoFe-LDH powder, which subsequently became mixed with the catalyst powder. The post-generated multiphase (oxy)hydroxides contain numerous oxygen vacancy defects (Figure 1b), which may serve as catalytic active sites to promote the adsorption and conversion of reactants and thereby enhancing catalytic performance [27,28].
As shown in the scanning electron microscopy (SEM) images (Figure 2), the nickel foam (NF) surface is uniformly covered with dense flake-like clusters of NiCoFe-LDH catalyst. On the upper layer of these clusters, the surface morphology of catalyst exhibits a hydrangea-like shape at the micrometer scale. The formation of this unique surface morphology can be attributed to the natural crystallization of the NiCoFe-LDH catalyst on the surface of NF, which significantly increases the specific surface area of catalyst, thereby providing more active sites for catalytic reactions and enhancing mass-transfer efficiency.

3.2. Electrocatalytic Activity

This work systematically evaluated the electrocatalytic activity of the synthesized NiCoFe-LDH catalyst toward the SOR. It should be noted that S2− has a corrosive effect on the catalyst [29,30]. In this work, through potentiodynamic scanning tests, the potentiodynamic polarization curve of the NiCoFe-LDH electrode in an electrolyte solution of 1 M KOH + (0.1–0.5) M Na2S were obtained (Figure S1), and corrosion potential (Ecorr) and corrosion current density (Icorr) were analyzed (Table S2). Among them, the electrolyte containing 1 M KOH + 0.1 M Na2S exhibited the highest Ecorr (−0.649 V) and the lowest Icorr (1.53 × 10−3 A cm−2), indicating that it had the least corrosive effect on the NiCoFe-LDH electrode [31]. As a result, 0.1 M Na2S was determined as the typical concentration to be used in the subsequent SOR performance tests. LSV curves in 1 M KOH and in 1 M KOH with added 0.1 M Na2S demonstrate that the NiCoFe-LDH-catalyzed reaction achieves a higher current density at the same potential, indicating significantly enhanced OER and SOR catalytic performance compared to commercial nickel foam (NF) and RuO2 (Figure 3a). At current densities of 10 mA cm−2, 50 mA cm−2, and 100 mA cm−2, the SOR driving potential for NiCoFe-LDH is only 0.381 V, 0.595 V, and 0.7 V vs. RHE, respectively. These values are markedly lower than those for NF and RuO2 (Figure 3b). Tafel slope is a key kinetic parameter for evaluating electrochemical catalytic performance and plays a crucial role in assessing catalyst activity. It provides essential insights for studying the kinetics of the SOR. A lower Tafel slope value indicates reaction kinetics under the action of catalyst, attributed to optimized reaction pathway and improved charge-transfer efficiency [32]. The catalytic kinetics of the SOR were systematically studied by deriving and analyzing Tafel slopes based on LSV curves. As shown in Figure 3c, the Tafel slope for NiCoFe-LDH is 91.2 mV dec−1, lower than those of commercial nickel foam (NF) (118.2 mV dec−1) and RuO2 (157.5 mV dec−1). This demonstrates the superiority of NiCoFe-LDH in the reaction kinetics of the SOR. To contrast with the conventional OER, the LSV curves of NiCoFe-LDH in 1 M KOH + 0.1 M Na2S were compared with those in 1 M KOH without Na2S (Figure 3d,e). It can be seen that in the presence of 0.1 M Na2S, the potentials at current densities of 10, 50, and 100 mA cm−2 are 0.98 V, 0.84 V, and 0.85 V (vs. RHE) lower, respectively, than those required for conventional OER.
To obtain a quantitative analysis of the catalyst kinetics, the equivalent circuit model shown in Figure 3f was used to fit the electrochemical impedance spectroscopy (EIS) data. The fitting results revealed that the NiCoFe-LDH electrode exhibits the smallest semicircle radius and the lowest charge-transfer resistance (Rct) of 0.4704 Ω (Table S3), compared to NF and RuO2. This indicates that NiCoFe-LDH possesses a faster charge-transfer rate and higher electrical conductivity [19]. To gain deeper insight into the superior SOR catalytic activity of NiCoFe-LDH, the double-layer capacitance (Cdl) derived from Figure 3g was used to estimate the electrochemically active surface area (ECSA) of catalyst. Cyclic voltammetry (CV) measurements at various scan rates (5–100 mV s−1, Figure S2) were conducted within a non-Faradaic potential window (OCP ± 0.05 V vs. RHE) to avoid interference from redox reactions. The Cdl values for NF and NiCoFe-LDH, calculated from the linear slope of the plot of current density difference at the open-circuit potential (Δj/2) versus scan rate, were determined to be 0.42 mF cm−2 and 3.21 mF cm−2, respectively. The corresponding ECSA values, calculated according to Equation (3), are 10.5 cm−2 and 80.25 cm−2, respectively. It is evident that the unique structure of hydrangea-like microspheres stacked on flake-like clusters endows NiCoFe-LDH with a significantly larger catalytic surface area and more active sites compared to bare NF. This explains its outstanding SOR catalytic performance. Compared with other reported catalysts, NiCoFe-LDH achieves excellent catalytic performance at lower concentrations of Na2S, whereas other catalysts rely on higher Na2S concentrations to enhance the reaction rate and reduce the required potential (Figure 3h, Table S4) [33,34,35,36,37,38,39,40]. This indicates that the SOR catalytic activity of NiCoFe-LDH surpasses that of most previously reported catalysts. Furthermore, in this work, the pH of the electrolyte without Na2S addition, with Na2S addition, and after the SOR were measured using the pH meter. The results showed that there were no significant differences among these three cases (Figure S3), thus eliminating the interference of pH changes on the performance evaluation of SOR.
To assess the mechanical and chemical stability of NiCoFe-LDH, a chronopotentiometry (CP) measurement was conducted at a constant current density of 100 mA cm−2 for 100 h (Figure 3i). The results demonstrate outstanding stability, with a negligible potential fluctuation of only 40 mV during prolonged operation. After stability testing, the NiCoFe-LDH catalyst grown on the surface of NF underwent in situ transformation into a colony-like structure firmly anchored to the substrate (Figure 4). This robust structure ensures the continuous provision of a large active surface area during catalytic reactions, facilitating efficient mass transport, which is proved by the extremely minor potential changes (23 mV at 100 mA cm−2) before and after the stability test (inset of Figure 3i).
Leveraging the outstanding SOR catalytic activity of the NiCoFe-LDH catalyst, an AEM electrolyzer was employed to construct an SOR-coupled HER electrolysis system, as illustrated in Figure 5a,b. In this system, NiCoFe-LDH and NF served as the anode and cathode, respectively (both with an electrode area of 2 cm × 2 cm), to evaluate the performance of SOR-assisted water electrolysis for hydrogen production. The testing of this system was conducted under simulated industrial operating conditions (6 M KOH + 0.1 M Na2S, 80 °C). The S2− is oxidized to elemental sulfur (S) at the anode of the AEM electrolyzer, catalyzed by NiCoFe-LDH. The generated S subsequently reacts with excess S2− in the electrolyte to form soluble polysulfides (Sn2−), thereby achieving S2− degradation, while green hydrogen is simultaneously produced at the cathode (Figure 5a). As shown in the inset of Figure 5b, the anode electrolyte gradually turned yellow, indicating the progressive accumulation of Sn2− during the SOR process [41,42,43]. Figure 5c presents the LSV curves for the NiCoFe-LDH||NF electrode pair in both the SOR/HER electrolysis system and the conventional OER/HER electrolysis system. It is evident that at the same applied cell voltage, the SOR/HER system achieves significantly higher current densities than the traditional OER/HER system. The cell potentials for the SOR-coupled HER system at current densities of 100 and 200 mA cm−2 are only 0.72 V and 0.85 V, respectively, representing reductions of 1.05 V and 1.07 V compared to the conventional water electrolysis system (Figure 5d). Consequently, the corresponding electrical energy consumption is reduced by 59.3% and 55.7%, respectively. This significant energy saving is primarily attributed to the faster kinetics of the anodic SOR compared to OER, highlighting the crucial role of SOR in lowering the electrolysis potential required for water splitting. The above information demonstrates that the NiCoFe-LDH-catalyzed SOR-coupled HER electrolysis system achieves the objective of energy-efficient hydrogen production.
To evaluate the hydrogen production efficiency of this system, electrolysis was conducted at a constant current of 100 mA (current density of 25 mA cm−2) for one hour. The hydrogen gas produced was collected and measured via the water-displacement method. The experimentally measured hydrogen yield was nearly identical to the theoretical value (Figure 5e). The Faradaic efficiency (FE) was calculated to be as high as 95.67% based on the ratio of the actual yield to the theoretical yield. The energy consumption for hydrogen production by this electrolysis system, calculated from the electrical energy consumed during electrolysis and the measured actual hydrogen yield, was determined to be 1.42 kWh m−3 H2 (only for electrolysis, excluding losses such as pumping, gas separation and system heat loss). This is significantly lower than the theoretical minimum energy consumption (2.94 kWh m−3 H2) for OER/HER electrolysis systems [9]. To verify its feasibility under harsh industrial operating conditions, a durability test was performed at a current density of 100 mA cm−2. The NiCoFe-LDH catalyst maintained stable operation for 100 h at 100 mA cm−2 with negligible voltage fluctuation (Figure 5f). This indicates its ability to meet the stability requirements for industrial-scale applications, demonstrating its strong potential for industrial-scale deployment. In Figure 5g, the mechanistic diagram of SOR/HER system is illustrated, which clarifies the promoting effect of SOR on the alkaline water electrolysis.

3.3. Theoretical Analysis

From the above experiments, it can be found that multiphase NiCoFe-LDH mainly consists of Fe2O3, CoOOH, and Ni(OH)2, which may all make contributions to electrocatalytic SOR. From Figure 6a–c, it can be seen that the total energies of the three slab models are different from each other, which could explain the distinct adsorption energies of *S on them. Through accurate calculation, the adsorption energies of S onto Fe2O3, CoOOH and Ni(OH)2 are all negative values, proving that the *S adsorption on them are all exothermic processes and spontaneous. After a comparison of the corresponding *S adsorption intensities, the following rules can be established: CoOOH > Ni(OH)2 > Fe2O3, which attests that CoOOH is more active for SOR than the other two.

4. Conclusions

In summary, this study synthesized a multiphase NiCoFe-based LDH (namely NiCoFe-LDH) via a facile seed-assisted heterogeneous nucleation approach. The NiCoFe-LDH possesses a unique micro-sized hydrangea-like morphology with an extremely large specific surface area and synergistically active phases, providing abundant catalytic interfaces and reactive centers for the SOR. Consequently, the NiCoFe-LDH catalytic electrode achieved an ultralow electrolysis potential of merely 0.381 V at 10 mA cm−2 in 1 M KOH + 0.1 M Na2S electrolyte, significantly reducing the potential by 0.98 V compared to the traditional OER. Under stringent industrial operating conditions (6 M KOH + 0.1 M Na2S, 80 °C), the electrolysis potential for the SOR/HER system based on the NiCoFe-LDH||NF electrode pair was reduced to 0.71 V at 100 mA cm−2, demonstrating a substantial reduction of 1.05 V compared to the conventional OER/HER system. Furthermore, the NiCoFe-LDH||NF electrode pair maintained stable hydrogen production and S2− oxidation or degradation for 100 h with negligible potential fluctuation, while the energy consumption for hydrogen production was lowered to 1.42 kWh m−3 H2. From a theoretical perspective, the adsorptions of S onto Fe2O3, CoOOH, and Ni(OH)2 are all spontaneous, and the corresponding intensities are as follows: CoOOH > Ni(OH)2 > Fe2O3. Evidently, the outstanding SOR catalytic performance of NiCoFe-LDH renders it highly promising for achieving industrial-scale energy-efficient production of green hydrogen.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/ma18143377/s1, Figure S1: Potentiodynamic polarization curve of NiCoFe-LDH in 1 M KOH + (0.1–0.5) M Na2S; Figure S2: CV curves (OCP ± 0.05 V vs. RHE) of (a) NiCoFe-LDH and (b) NF electrodes under different scan rates (5–100 mV s−1); Figure S3: The pH of the electrolyte without Na2S addition, with Na2S addition, and after the SOR. Table S1: A comparison of the electro-oxidation potentials of the SOR with those of other electrolyte oxidation reactions; Table S2: Corrosion potential (Ecorr) and corrosion current density (Icorr) of NICoFe-LDH in 1 M KOH + (0.1–0.5) M Na2S; Table S3: Electrochemical impedance parameters obtained by fitting the Nyquist plots of NiCoFe-LDH, RuO2, and NF to the equivalent circuit mode; Table S4: A comparison of the electrochemical SOR activities of this work with recent outstanding reported electrocatalysts.

Author Contributions

Conceptualization, M.W.; Methodology, Z.L. and Y.N.; Formal analysis, Z.L. and Y.N.; Investigation, Z.L. and H.D.; Writing—original draft preparation, Z.L.; Writing—review and editing, Z.L., H.D. and M.W.; Visualization, Z.L. and Y.N.; Supervision, P.L. and G.M. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by Central Guidance Fund for Local Science and Technology Development (YDZJSX2024D035), Fundamental Research Program of Shanxi Province (202303021222124) and Postdoctoral Fellowship Program of CPSF (GZC20232457).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The original contributions presented in this study are included in the article/Supplementary Materials. Further inquiries can be directed to the corresponding authors.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Qiao, Y.; Peng, M.; Lan, J.; Jiang, K.; Chen, D.; Tan, Y. Active-site engineering in dealloyed nanoporous catalysts for electrocatalytic water splitting. J. Mater. Chem. A 2022, 11, 495–511. [Google Scholar] [CrossRef]
  2. Zhou, Q.; Hao, Q.; Li, Y.; Yu, J.; Xu, C.; Liu, H.; Yan, S. Free-standing trimodal porous NiZn intermetallic and Ni heterojunction as highly efficient hydrogen evolution electrocatalyst in the alkaline electrolyte. Nano Energy 2021, 89, 106402. [Google Scholar] [CrossRef]
  3. Bu, X.; Liang, X.; Bu, Y.; Quan, Q.; Meng, Y.; Lai, Z.; Wang, W.; Liu, C.; Lu, J.; Wu, C.-M.L.; et al. NiMo@C3N5 heterostructures with multiple electronic transmission channels for highly efficient hydrogen evolution from alkaline electrolytes and seawater. Chem. Eng. J. 2022, 438, 135379. [Google Scholar] [CrossRef]
  4. Karthikeyan, S.C.; Ramakrishnan, S.; Prabhakaran, S.; Subramaniam, M.R.; Mamlouk, M.; Kim, D.H.; Yoo, D.J. Low-Cost Self-Reconstructed High Entropy Oxide as an Ultra-Durable OER Electrocatalyst for Anion Exchange Membrane Water Electrolyzer. Small 2024, 20, 2402241. [Google Scholar] [CrossRef]
  5. Zhang, M.; Cui, H.; Ji, L.; Zhang, J.; Wang, C.; Guan, R. Durable Hierarchical Porous PtCo Nanoparticle Films for Hydrogen Evolution Reaction in Simulated Seawater. ACS Appl. Nano Mater. 2025, 8, 8176–8186. [Google Scholar] [CrossRef]
  6. Zhu, Y.; Li, L.; Cheng, H.; Ma, J. Alkaline Hydrogen Evolution Reaction Electrocatalysts for Anion Exchange Membrane Water Electrolyzers: Progress and Perspective. JACS Au 2024, 4, 4639–4654. [Google Scholar] [CrossRef] [PubMed]
  7. Chen, Y.; Rong, J.; Fan, Q.; Sun, M.; Deng, Q.; Ni, Z.; Li, X.; Hu, T. Facile engineering of CoS/rGO heterostructures on carbon cloth for efficient all-pH hydrogen evolution reaction and alkaline water electrolysis. J. Mater. Chem. A 2024, 13, 486–498. [Google Scholar] [CrossRef]
  8. Liang, Z.; Shen, D.; Wei, Y.; Sun, F.; Xie, Y.; Wang, L.; Fu, H. Modulating the Electronic Structure of Cobalt-Vanadium Bimetal Catalysts for High-Stable Anion Exchange Membrane Water Electrolyzer. Adv. Mater. 2024, 36, 2408634. [Google Scholar] [CrossRef] [PubMed]
  9. Wang, J.; Zhou, M.; Fu, R.; Ge, J.; Yang, W.; Hong, X.; Sun, C.; Liao, X.; Zhao, Y.; Wang, Z. Iron Molybdenum Sulfide-Supported Ultrafine Ru Nanoclusters for Robust Sulfion Degradation-Assisted Hydrogen Production. Adv. Funct. Mater. 2024, 34, 2315326. [Google Scholar] [CrossRef]
  10. Lyu, C.; Li, Y.; Cheng, J.; Yang, Y.; Wu, K.; Wu, J.; Wang, H.; Lau, W.-M.; Tian, Z.; Wang, N.; et al. Dual Atoms (Fe, F) Co-Doping Inducing Electronic Structure Modulation of NiO Hollow Flower-Spheres for Enhanced Oxygen Evolution/Sulfion Oxidation Reaction Performance. Small 2023, 19, 2302055. [Google Scholar] [CrossRef] [PubMed]
  11. Huang, C.-J.; Xu, H.-M.; Shuai, T.-Y.; Zhan, Q.-N.; Zhang, Z.-J.; Li, G.-R. Modulation Strategies for the Preparation of High-Performance Catalysts for Urea Oxidation Reaction and Their Applications. Small 2023, 19, 2301130. [Google Scholar] [CrossRef] [PubMed]
  12. Wang, T.; Miao, L.; Zheng, S.; Qin, H.; Cao, X.; Yang, L.; Jiao, L. Interfacial Engineering of Ni3N/Mo2N Heterojunctions for Urea-Assisted Hydrogen Evolution Reaction. ACS Catal. 2023, 13, 4091–4100. [Google Scholar] [CrossRef]
  13. Liu, W.-J.; Xu, Z.; Zhao, D.; Pan, X.-Q.; Li, H.-C.; Hu, X.; Fan, Z.-Y.; Wang, W.-K.; Zhao, G.-H.; Jin, S.; et al. Efficient electrochemical production of glucaric acid and H2 via glucose electrolysis. Nat. Commun. 2020, 11, 265. [Google Scholar] [CrossRef] [PubMed]
  14. Shi, H.; Wang, T.Y.; Lin, Z.J.; Liu, S.X.; Liu, X.; Zhou, R.X.; Cai, Z.; Huang, Y.H.; Li, Q. Spontaneous Hydrogen Production Coupled with Glucose Valorization through Modulating Au-Pt Coordination on Ultrathin Au3Pt Twin Nanowires. Angew. Chem. Int. Ed. 2025, 64, e202424476. [Google Scholar] [CrossRef]
  15. Pei, A.; Xie, R.K.; Zhu, L.H.; Wu, F.S.; Huang, Z.A.; Pang, Y.Y.; Chang, Y.C.; Chai, G.L.; Pao, C.W.; Gao, Q.S.; et al. Methanol-Enhanced Low-Cell-Voltage Hydrogen Generation at Industrial-Grade Current Density by Triadic Active Sites of Pt1-Pdn-(Ni,Co)(OH)x. J. Am. Chem. Soc. 2025, 147, 3185–3194. [Google Scholar] [CrossRef] [PubMed]
  16. Xu, K.; Liang, L.; Li, T.; Bao, M.; Yu, Z.; Wang, J.; Thalluri, S.M.; Lin, F.; Liu, Q.; Cui, Z.; et al. Pt1.8Pd0.2CuGa Intermetallic Nanocatalysts with Enhanced Methanol Oxidation Performance for Efficient Hybrid Seawater Electrolysis. Adv. Mater. 2024, 36, 2403792. [Google Scholar] [CrossRef] [PubMed]
  17. Feng, C.; Lv, M.; Shao, J.; Wu, H.; Zhou, W.; Qi, S.; Deng, C.; Chai, X.; Yang, H.; Hu, Q.; et al. Lattice Strain Engineering of Ni2P Enables Efficient Catalytic Hydrazine Oxidation-Assisted Hydrogen Production. Adv. Mater. 2023, 35, 2305598. [Google Scholar] [CrossRef] [PubMed]
  18. Huang, J.-L.; Zhang, H.; Suo, T.-Y.; Cunha, J.; Yu, Z.-P.; Xu, W.-Y.; Chen, L.; Hou, Z.-H.; Yin, H. Ru nanoclusters immobilized in N-doped porous carbon for efficient hydrazine-assisted hydrogen production and Zn–hydrazine battery. Rare Met. 2025, 44, 2502–2512. [Google Scholar] [CrossRef]
  19. Yu, Z.; Boukhvalov, D.W.; Tan, H.; Xiong, D.; Feng, C.; Wang, J.; Wang, W.; Zhao, Y.; Xu, K.; Su, W.; et al. Sulfur and phosphorus co-doped FeCoNiCrMn high-entropy alloys as efficient sulfion oxidation reaction catalysts enabling self-powered asymmetric seawater electrolysis. Chem. Eng. J. 2024, 494, 153094. [Google Scholar] [CrossRef]
  20. Zhu, Y.; Wang, S.; Chen, Y.; Zhang, Y.; Feng, Y.; Zhang, G. Screened d-p Orbital Hybridization in Turing Structure of Confined Nickel for Sulfion Oxidation Accelerated Hydrogen Production. Angew. Chem. Int. Ed. 2024, 64, e202419572. [Google Scholar] [CrossRef] [PubMed]
  21. Wang, P.; Zhang, X.; Zhou, B.; Meng, F.; Wang, Y.; Wen, G. Recent advance of layered double hydroxides materials: Structure, properties, synthesis, modification and applications of wastewater treatment. J. Environ. Chem. Eng. 2023, 11, 111191. [Google Scholar] [CrossRef]
  22. Zhai, Y.; Ren, X.; Gan, T.; She, L.; Guo, Q.; Yang, N.; Wang, B.; Yao, Y.; Liu, S. Deciphering the Synergy of Multiple Vacancies in High-Entropy Layered Double Hydroxides for Efficient Oxygen Electrocatalysis. Adv. Energy Mater. 2025, 2502065. [Google Scholar] [CrossRef]
  23. Zhai, P.; Xia, M.; Wu, Y.; Zhang, G.; Gao, J.; Zhang, B.; Cao, S.; Zhang, Y.; Li, Z.; Fan, Z.; et al. Engineering single-atomic ruthenium catalytic sites on defective nickel-iron layered double hydroxide for overall water splitting. Nat. Commun. 2021, 12, 4587. [Google Scholar] [CrossRef] [PubMed]
  24. Ai, L.; Tian, Y.; Xiao, T.; Zhang, J.; Zhang, C.; Jiang, J. Energy-saving hydrogen production from sulfion oxidation-hybrid seawater splitting enabled by superwettable corrosion-resistant NiFe layered double hydroxide/FeNi2S4 heterostructured nanoarrays. J. Colloid Interface Sci. 2024, 673, 607–615. [Google Scholar] [CrossRef] [PubMed]
  25. Jiang, J.; Tian, Y.; Zhang, J.; Zhang, C.; Ai, L. Metallic Cu-incorporated NiFe layered double hydroxide nanosheets enabling energy-saving hydrogen generation from chlorine-free seawater electrolysis coupled with sulfion upcycling. Fuel 2024, 367, 131506. [Google Scholar] [CrossRef]
  26. Li, Z.; Lin, G.; Wang, L.; Lee, H.; Du, J.; Tang, T.; Ding, G.; Ren, R.; Li, W.; Cao, X.; et al. Seed-assisted formation of NiFe anode catalysts for anion exchange membrane water electrolysis at industrial-scale current density. Nat. Catal. 2024, 7, 944–952. [Google Scholar] [CrossRef]
  27. Xie, C.; Chen, W.; Wang, Y.; Yang, Y.; Wang, S. Dynamic evolution processes in electrocatalysis: Structure evolution, characterization and regulation. Chem. Soc. Rev. 2024, 53, 10852–10877. [Google Scholar] [CrossRef] [PubMed]
  28. Zhao, J.-W.; Wang, H.-Y.; Feng, L.; Zhu, J.-Z.; Liu, J.-X.; Li, W.-X. Crystal-Phase Engineering in Heterogeneous Catalysis. Chem. Rev. 2023, 124, 164–209. [Google Scholar] [CrossRef] [PubMed]
  29. Wu, H.; Wang, C.; Ma, Y.; Huang, S.; Ren, Y.; Ding, F.; Li, F.; Yang, Y.; Gu, J.; Tang, S.; et al. NiO/RuO2p-n Heterojunction Nanofoam as a High-Performance Electrocatalyst for Desulfurization and Concurrent Hydrogen Evolution. Lnorg. Chem. 2024, 63, 12604–12614. [Google Scholar] [CrossRef] [PubMed]
  30. Yu, H.; Wang, W.; Mao, Q.; Deng, K.; Xu, Y.; Wang, Z.; Li, X.; Wang, H.; Wang, L. Electrocatalytic sulfion recycling assisted energy-saving hydrogen production using CuCo-based nanosheet arrays. J. Mater. Chem. A 2023, 11, 2218–2224. [Google Scholar] [CrossRef]
  31. Hou, Z.; Liu, H.; Li, W.; Chen, C.; Zhang, Z.; Dong, J.; Lu, W.; Feng, Q.; Jia, B.; Song, K. Effect of trace Sc on corrosion behavior of titanium material in cathode roller for electrolytic copper foil. Corros. Sci. 2025, 246, 112722. [Google Scholar] [CrossRef]
  32. Zhang, H.; Cheng, C.; Zhou, J.; Ma, C.; Shi, P.; Wu, H.; Yin, P.; Cao, W.; Xia, J.; Zhu, L.; et al. Atomically dispersed ruthenium sites with electron-rich environments in intermetallic compounds for high-current-density hydrogen evolution. J. Mater. Chem. A 2023, 11, 10328–10336. [Google Scholar] [CrossRef]
  33. Kim, K.; Han, J.-I. Carbon-supported bimetallic Pd–Ir catalysts for alkaline sulfide oxidation in direct alkaline sulfide fuel cell. J. Appl. Electrochem. 2015, 45, 533–539. [Google Scholar] [CrossRef]
  34. Kim, K.; Han, J.-I. Carbon supported bimetallic Pd–Co catalysts for alkaline sulfide oxidation in direct alkaline sulfide fuel cell. Int. J. Hydrogen Energy 2015, 40, 4567–4572. [Google Scholar] [CrossRef]
  35. Kim, K.; Son, J.; Han, J.-I. Metal sulfides as anode catalysts in direct alkaline sulfide fuel cell. Int. J. Hydrogen Energy 2014, 39, 10493–10497. [Google Scholar] [CrossRef]
  36. Sahoo, P.C.; Kim, K.; Lee, J.H.; Han, J.-I.; Oh, Y.-K. Biomimetically Synthesized Hierarchical TiO2-Graphitic Carbon as Anodic Catalysts for Direct Alkaline Sulfide Fuel Cell. ACS Sustain. Chem. Eng. 2015, 3, 1764–1770. [Google Scholar] [CrossRef]
  37. Wang, Z.; Yang, G.; Tian, P.; Li, X.; Deng, K.; Yu, H.; Xu, Y.; Wang, H.; Wang, L. Hydrophilic functionalization of rhodium metallene for saving-energy hydrogen production and sulfur recovery. Chem. Eng. J. 2023, 473, 145147. [Google Scholar] [CrossRef]
  38. Yi, L.; Ji, Y.; Shao, P.; Chen, J.; Li, J.; Li, H.; Chen, K.; Peng, X.; Wen, Z. Scalable Synthesis of Tungsten Disulfide Nanosheets for Alkali-Acid Electrocatalytic Sulfion Recycling and H2 Generation. Angew. Chem. Int. Ed. 2021, 60, 21550–21557. [Google Scholar] [CrossRef] [PubMed]
  39. Li, Y.; Duan, Y.; Zhang, K.; Yu, W. Efficient anodic chemical conversion to boost hydrogen evolution with low energy consumption over cobalt-doped nickel sulfide electrocatalyst. Chem. Eng. J. 2022, 433, 134472. [Google Scholar] [CrossRef]
  40. Zhang, S.; Zhou, Q.; Shen, Z.; Jin, X.; Zhang, Y.; Shi, M.; Zhou, J.; Liu, J.; Lu, Z.; Zhou, Y.N.; et al. Sulfophobic and Vacancy Design Enables Self-Cleaning Electrodes for Efficient Desulfurization and Concurrent Hydrogen Evolution with Low Energy Consumption. Adv. Funct. Mater. 2021, 31, 2101922. [Google Scholar] [CrossRef]
  41. Ding, J.; Zhang, L.; Wei, Z.; Wang, Z.; Liu, Q.; Hu, G.; Luo, J.; Liu, X. Coupling Nitrate-to-Ammonia Conversion and Sulfion Oxidation Reaction Over Hierarchical Porous Spinel MFe2O4 (M = Ni, Co, Fe, Mn) in Wastewater. Small 2025, 21, 2411317. [Google Scholar] [CrossRef] [PubMed]
  42. Tan, H.; Yu, Z.; Zhang, C.; Lin, F.; Ma, S.; Huang, H.; Li, H.; Xiong, D.; Liu, L. Self-supported NiTe@NiMo electrodes enabling efficient sulfion oxidation reaction toward energy-saving and chlorine-free hybrid seawater electrolysis at high current densities. Energy Environ. Sci. 2024, 18, 1440–1451. [Google Scholar] [CrossRef]
  43. Yang, M.; Ding, J.; Wang, Z.; Zhang, J.; Peng, Z.; Liu, X. NiMo-based alloy and its sulfides for energy-saving hydrogen production via sulfion oxidation assisted alkaline seawater splitting. Chin. Chem. Lett. 2025, 25, 110861. [Google Scholar] [CrossRef]
Scheme 1. Synthesis process of NiCoFe-LDH catalyst.
Scheme 1. Synthesis process of NiCoFe-LDH catalyst.
Materials 18 03377 sch001
Figure 1. (a) XRD pattern and (b) EPR spectra of NiCoFe-LDH.
Figure 1. (a) XRD pattern and (b) EPR spectra of NiCoFe-LDH.
Materials 18 03377 g001
Figure 2. (ad) Low- and high-magnification SEM images of NiCoFe-LDH.
Figure 2. (ad) Low- and high-magnification SEM images of NiCoFe-LDH.
Materials 18 03377 g002
Figure 3. (a) LSV curves in 1 M KOH and 1 M KOH + 0.1 M Na2S; (b) comparison of SOR performances at 10, 50 and 100 mA cm−2; (c) Tafel plots; (d,e) comparison of OER and SOR of NiCoFe-LDH; (f) Nyquist plots; (g) Cdl values; (h) SOR performance contrast at 10 mA cm−2; (i) CP profile at 100 mA cm−2—the inset shows LSV curves before and after stability test.
Figure 3. (a) LSV curves in 1 M KOH and 1 M KOH + 0.1 M Na2S; (b) comparison of SOR performances at 10, 50 and 100 mA cm−2; (c) Tafel plots; (d,e) comparison of OER and SOR of NiCoFe-LDH; (f) Nyquist plots; (g) Cdl values; (h) SOR performance contrast at 10 mA cm−2; (i) CP profile at 100 mA cm−2—the inset shows LSV curves before and after stability test.
Materials 18 03377 g003
Figure 4. (ad) SEM images of NiCoFe-LDH after 100 h stability test in 1 M KOH + 0.1 M Na2S electrolyte.
Figure 4. (ad) SEM images of NiCoFe-LDH after 100 h stability test in 1 M KOH + 0.1 M Na2S electrolyte.
Materials 18 03377 g004
Figure 5. (a) A schematic diagram and (b) photograph of SOR/HER electrolysis system based on NiCoFe-LDH||NF pair—the inset shows the change in electrolyte color as the SOR progresses; (c) LSV curves of SOR/HER and OER/HER electrolysis system; (d) a comparison of cell voltages at 10–250 mA cm−2; (e) the amount of the theoretically calculated and experimentally measured H2 production—the inset is an illustration of H2 yields at 100 mA for 60 min; (f) CP profiles at 100 mA cm−2; (g) a mechanistic diagram for the effect of SOR on HER.
Figure 5. (a) A schematic diagram and (b) photograph of SOR/HER electrolysis system based on NiCoFe-LDH||NF pair—the inset shows the change in electrolyte color as the SOR progresses; (c) LSV curves of SOR/HER and OER/HER electrolysis system; (d) a comparison of cell voltages at 10–250 mA cm−2; (e) the amount of the theoretically calculated and experimentally measured H2 production—the inset is an illustration of H2 yields at 100 mA for 60 min; (f) CP profiles at 100 mA cm−2; (g) a mechanistic diagram for the effect of SOR on HER.
Materials 18 03377 g005
Figure 6. The slab models and corresponding total energies of (a) Fe2O3, (b) CoOOH, and (c) Ni(OH)2. (df) The top view of optimized *S adsorption models. (gi) A side view and corresponding adsorption energies of optimized *S adsorption models.
Figure 6. The slab models and corresponding total energies of (a) Fe2O3, (b) CoOOH, and (c) Ni(OH)2. (df) The top view of optimized *S adsorption models. (gi) A side view and corresponding adsorption energies of optimized *S adsorption models.
Materials 18 03377 g006
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Liang, Z.; Nian, Y.; Du, H.; Li, P.; Wang, M.; Ma, G. Multiphase NiCoFe-Based LDH for Electrocatalytic Sulfion Oxidation Reaction Assisting Efficient Hydrogen Production. Materials 2025, 18, 3377. https://doi.org/10.3390/ma18143377

AMA Style

Liang Z, Nian Y, Du H, Li P, Wang M, Ma G. Multiphase NiCoFe-Based LDH for Electrocatalytic Sulfion Oxidation Reaction Assisting Efficient Hydrogen Production. Materials. 2025; 18(14):3377. https://doi.org/10.3390/ma18143377

Chicago/Turabian Style

Liang, Zengren, Yong Nian, Hao Du, Peng Li, Mei Wang, and Guanshui Ma. 2025. "Multiphase NiCoFe-Based LDH for Electrocatalytic Sulfion Oxidation Reaction Assisting Efficient Hydrogen Production" Materials 18, no. 14: 3377. https://doi.org/10.3390/ma18143377

APA Style

Liang, Z., Nian, Y., Du, H., Li, P., Wang, M., & Ma, G. (2025). Multiphase NiCoFe-Based LDH for Electrocatalytic Sulfion Oxidation Reaction Assisting Efficient Hydrogen Production. Materials, 18(14), 3377. https://doi.org/10.3390/ma18143377

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop