Next Article in Journal
Elucidating the Synergism by Applying Ni-Cu/Cr2O3 Catalysts for Green Methanol Fuel Synthesis by CO2 Hydrogenation
Next Article in Special Issue
Enhanced Oxygen Reduction Reaction Activity of Carbon-Supported Pt-Co Catalysts Prepared by Electroless Deposition and Galvanic Replacement
Previous Article in Journal
Citrus aurantiifolia Peel-Facilitated Synthesis of Zinc Oxide, Interfaced with Biomass-Assisted Graphene Oxide for Enhanced Photocatalytic Degradation of Dye
Previous Article in Special Issue
Characteristic Influence of Cerium Ratio on PrMn Perovskite-Based Cathodes for Solid Oxide Fuel Cells
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Iron-Doped NiSe2 and Its Enhanced Oxygen Evolution Reaction Activity

1
School of Electronic Information and Electrical Engineering, Chengdu University, Chengdu 610106, China
2
School of Physics, University of Electronic Science and Technology of China, Chengdu 610054, China
3
Yangtze Delta Region Institute (Huzhou), University of Electronic Science and Technology of China, Huzhou 313001, China
*
Author to whom correspondence should be addressed.
Catalysts 2025, 15(9), 876; https://doi.org/10.3390/catal15090876
Submission received: 15 August 2025 / Revised: 4 September 2025 / Accepted: 8 September 2025 / Published: 12 September 2025

Abstract

Doping a third element or external functional components into binary alloy nanostructured catalysts typically significantly enhances their electrocatalytic performance. This study demonstrates that doping nickel selenide (NiSe2) with approximately 10   a t % iron (Fe) is an effective strategy for improving its oxygen evolution reaction (OER) catalytic activity. The resulting Ni0.9Fe0.1Se2 undergoes a structural transformation from its original nanodendritic morphology and exhibits outstanding OER catalytic performance in alkaline media. It achieves a low overpotential of 231   m V at a current density of 10   m A   c m 2 , which is approximately 30 % lower than that of NiSe2 ( 301   m V ). The Tafel slope of Ni0.9Fe0.1Se2 is 116   m V   d e c 1 . However, degradation observed after 5 h of stability testing suggests that the doping process requires further optimization.

1. Introduction

With the ever-increasing global demand for energy, coupled with the finite reserves of fossil fuels and their tendency to cause environmental pollution, the search for clean and renewable energy sources has become paramount. Among these, hydrogen energy stands out as a quintessential clean energy source, attracting significant attention. Hydrogen evolution reaction (HER) via water electrolysis offers advantages such as process simplicity, environmental friendliness, and recyclability [1,2]. The water splitting reaction comprises two half-reactions: The hydrogen evolution reaction (HER) at the cathode and the oxygen evolution reaction (OER) at the anode. Both reactions require highly efficient catalysts to achieve high current densities at low overpotentials [3,4]. However, in the water electrolysis process, OER, due to its sluggish kinetics, constitutes the rate-limiting step and typically necessitates substantial amounts of precious metal catalysts to reach the required current density ( > 10   m A / c m 2 ) [5]. Consequently, researchers have focused on exploring earth-abundant 3d transition metal compounds to develop low-cost, long-lasting, and highly active OER catalysts, such as perovskite oxides [6], metal phosphates [7,8], selenide composites [9,10], nickel borate composites [11,12], among others.
Among these 3d transition metal catalysts, transition metal compounds (e.g., oxides [13,14,15], phosphides [7,16], selenides [17,18]) are regarded as promising alternatives to precious metal OER catalysts, owing to their high catalytic activity, excellent stability, unique redox characteristics, and low cost [19,20,21]. In particular, metal selenides have garnered considerable interest in the OER field due to their tunable electronic structure, corrosion resistance, and favorable intrinsic activity [22,23]. For instance, Nath’s group reported Ni3Se2 synthesized via electrodeposition, requiring only 290   m V overpotential to achieve 10   m A   c m 2 in 0.3   M KOH [23]. Cai et al. demonstrated high stability for dendrite-like nanostructured NiSe2 synthesized via a co-reduction method, showing an overpotential decay of only ~ 2   m V after 10,000 cycles [24]. Chen et al. grew NiSe2 on nickel foam in a tube furnace at 500 ~ 600   ° C and used it as a highly efficient HER catalyst [25]. Furthermore, studies indicate that incorporating a third element or other external functional materials into binary alloy catalysts often significantly enhances their electrocatalytic performance. For example, research has found that even trace amounts of iron ( ~ 1   p p m ) doping can dramatically boost the OER activity of N i 2 + and C o 2 + based catalysts [26,27]. Boettcher and colleagues, while investigating the electrocatalytic water oxidation performance of solution-cast metal oxide films in alkaline medium, affirmed the enhancement effect of iron [28]. Subsequently, the same research group reported the effects of iron—either accidentally or intentionally incorporated into Ni(OH)2/NiOOH electrodes—during the OER in 1   M KOH. They also observed that the OER activity of Ni(OH)2 peaked when the iron content reached an atomic percentage between 20 % and 30 % [29,30]. As for iron doping in nickel selenides, it was first reported by Du et al., who synthesized dendritic nanostructured Ni0.5Fe0.5Se2 and urchin-like Ni1.12Fe0.45Se2 catalysts, both exhibiting outstanding OER activity and durability in alkaline media [31,32]. Later, Chang et al. synthesized Fe-doped nickel selenide on carbon nanotubes, which significantly enhanced their electrocatalytic oxygen evolution activity, requiring only an overpotential of 282.7   m V to achieve a current density of 10   m A   c m 2 [33]. Deng et al. fabricated an iron-doped nickel selenide (NF/NiSe/Ni3Se2-Fe) catalyst on nickel foam via electrodeposition. This catalyst exhibited overpotentials of 200   m V for OER at a current density of 10   m A   c m 2 [18].
It is evident that many studies have demonstrated that iron doping in transition metal compounds can significantly enhance the oxygen evolution reaction (OER) activity of nickel-based catalysts. For transition metal oxides, numerous systematic studies have reported the effects of iron doping on their morphology, composition, and catalytic performance, and the optimal iron doping levels have been comprehensively discussed. However, research on iron doping in transition metal selenides, particularly nickel selenides, remains insufficient. In particular, the influence of low-concentration iron doping on OER activity and the specific mechanisms underlying the activity enhancement have yet to be fully elucidated.
This study evaluates the performance of two NiFeSe catalysts with low iron doping levels for the oxygen evolution reaction (OER). The results demonstrate that Ni0.9Fe0.1Se2 exhibits the best OER catalytic activity among all tested catalysts. Its iron doping level is lower than those reported for Ni0.5Fe0.5Se2 and Ni1.12Fe0.45Se2 in the literature, yet it achieves an overpotential of only 231 mV at a current density of 10   m A   c m 2 , with a Tafel slope of 116   m V   d e c 1 .

2. Results and Discussion

This study successfully prepared NiSe2, FeSe2, and iron-doped nickel-iron selenide (Ni1−xFexSe2) nanomaterials using an efficient and controllable wet-chemical method. The key advantage of this approach lies in its ability to achieve molecular-level uniform dispersion and complete dissolution of reaction precursors in an organic solvent system under relatively low temperature and ambient pressure conditions, thereby providing an ideal environment for the nucleation and growth of nanocrystals with uniform morphology and controllable composition. Specifically, a mixed solution of oleylamine (OAm) and tetralin was selected as the reaction medium. Three reaction systems were designed to synthesize the target products by adjusting the precursor combinations: (1) In the single-nickel-source system, Ni(acac)2 and selenium powder (Se) were used as reactants to generate pure-phase NiSe2. (2) In the single-iron-source system, iron sulfate (Fe2(SO4)3) and selenium powder (Se) served as reactants to produce pure-phase FeSe2. (3) In the nickel-iron mixed-source system, Ni(acac)2, iron sulfate (Fe(SO4)3), and selenium powder (Se) were jointly used as reactants. In this system, both nickel and iron ions participated in the reaction in the solution to form Ni1−xFexSe2. The XRD patterns of the as-synthesized products are shown in Figure 1.
When the solute was Ni(acac)2 and selenium powder, the diffraction peaks exhibited the characteristic pattern of the cubic NiSe2 phase (pyrite-type structure). The diffraction peaks located at 2 θ values of 30.0 ° , 33.6 ° , 36.9 ° , 42.9 ° , 50.7 ° , 53.2 ° , 55.5 ° , and 57.8 ° correspond to the ( 200 ), ( 210 ), ( 211 ), ( 220 ), ( 311 ), ( 222 ), ( 023 ), and ( 321 ) crystal planes of cubic NiSe2 (JCPDS card No.88-1711), respectively. When the solute was Fe2(SO4)3 and selenium powder, the XRD pattern of the sample displayed the typical FeSe2 crystalline phase. Comparison with the standard JCPDS card No.65-1455 revealed that the diffraction peaks at 2 θ values of 29.3 ° , 31.1 ° , 34.8 ° , 36.2 ° , 48.2 ° , 50.9 ° , 51.0 ° , 53.9 ° , 57.5 ° , and 59.9 ° correspond to the ( 011 ), ( 101 ), ( 111 ), ( 120 ), ( 211 ), ( 002 ), ( 130 ), ( 031 ), ( 131 ), and ( 310 ) planes of FeSe2, respectively. Figure 1c compares the XRD patterns of samples prepared with different amounts of Fe2(SO4)3 with those of NiSe2 and FeSe2. The results indicate that the diffraction peak positions of samples prepared with 0.1   m m o l and 0.2   m m o l Fe2(SO4)3, respectively, are similar. Furthermore, these diffraction peaks correspond simultaneously to the characteristic peaks of FeSe2 (JCPDS No.65-1455) and NiSe2 (JCPDS No.88-1711). This demonstrates that the final product is the NixFe1−xSe2, signifying the coexistence of NiSe2 and FeSe2 phases. This finding is consistent with previous reports [32]. Furthermore, no detectable impurity peaks are observed in the pattern, suggesting the relatively high purity of the synthesized sample.
The morphology of the samples was characterized by SEM, as shown in Figure 2. The images reveal that undoped NiSe2 exhibits a nanodendritic structure, while FeSe2 displays a nanourchin-like morphology, which is consistent with the structures reported by Du et al. Upon iron doping, a coating layer gradually forms on the surface of NiSe2, resulting in the disruption of its original dendritic architecture. Although iron doping was achieved by adjusting the addition level of Fe2(SO4)3, the final composition of the samples required verification through characterization. For this purpose, elemental distribution was analyzed using the Energy Dispersive X-ray Spectroscopy (EDS) detector equipped on the SEM, obtaining quantitative compositional information. The EDS spectra of the samples obtained with the two different addition amounts are shown in Figure S1. Based on the EDS analysis, the atomic ratios of the various elements in the samples under different ferric sulfate doping levels were determined. When the amount of Fe2(SO4)3 added was 0.1   m m o l , the atomic percentages of Fe, Ni, and Se were 2.6 % , 23.4 % , and 62 % , respectively, corresponding to a chemical formula of Ni0.9Fe0.1Se2. When the addition amount was 0.2   m m o l , the atomic percentages of the three elements were 4.4 % , 21.7 % , and 60 % , respectively, corresponding to a chemical formula of Ni0.83Fe0.17Se2. It should be noted that due to the introduction of Fe2(SO4)3 during the preparation process, sulfate residues remained in the samples, resulting in signals of O and S in the EDS spectra. Furthermore, the atomic percentage of Se was significantly more than twice the sum of Fe and Ni, which is attributed to the addition of excess Se powder during the synthesis process. Additionally, from the SEM-EDS mapping results shown in Figure S2, it can be observed that the elements Fe, Ni, and Se are uniformly distributed in the sample.
To further investigate the effects of different Fe2(SO4)3 doping amounts on the products, we performed XPS characterization on the products, with the results shown in Figure 3. Figure 3a displays the XPS survey spectrum of the Ni0.9Fe0.1Se2 sample obtained by doping with 0.1   m m o l Fe2(SO4)3. Figure 3b–d present the corresponding high-resolution spectra of Fe2p, Ni2p, and Se3d for this sample, respectively. Figure 3e,f show the comparative spectra of Se3d and Ni2p for the four samples: Ni0.9Fe0.1Se2, Ni0.83Fe0.17Se2, FeSe2, and NiSe2. All spectra were calibrated using the C1s peak at the binding energy of 285.5   e V .
From the XPS survey spectrum in Figure 3a, characteristic peaks can be observed at binding energies of 854 eV (Ni 2p), 709 eV (Fe 2p), and 52 eV (Se 3d). Analysis of the Fe2p spectrum in Figure 3b reveals that the Fe2p3/2 peak can be deconvoluted into three peaks located at 707.6   e V , 709.3   e V , and 714.7   e V , attributed to the metallic iron, the iron component in NixFe1−xSe2, and its satellite peak, respectively [31,32]. Analysis of the Ni2p spectrum in Figure 3c indicates that for the Ni0.9Fe0.1Se2 sample, the peaks at 853.2   e V and 870.2   e V correspond to the Ni2p3/2 and Ni2p1/2 main peaks, respectively, while the peaks at 856.3   e V and 862.0   e V are assigned to nickel oxide species and the Ni satellite peak [31,32]. Analysis of the Se3d spectrum in Figure 3d shows peaks at 54.2   e V , 55.2   e V , and 58.8   e V , corresponding to Se3d5/2, Se3d3/2, and selenium oxide (SeOx), respectively [31,32]. The results in Figure 3a–d collectively confirm the successful incorporation of iron. Furthermore, comparative analysis of the Se3d and Ni2p spectra (Figure 3e,f) shows that the Se3d5/2 peak position ( 54.2   e V ) in Ni0.9Fe0.1Se2 and Ni0.83Fe0.17Se2 is lower than that in pure NiSe2 ( 54.4   e V ). Similarly, the Ni2p3/2 peak position ( 853.2   e V ) in Ni0.9Fe0.1Se2 and Ni0.83Fe0.17Se2 is lower than that in pure NiSe2 ( 853.5   e V ). In summary, compared to NiSe2, both Ni and Se binding energies in the doped samples (Ni0.9Fe0.1Se2, Ni0.83Fe0.17Se2) exhibit a shift towards lower binding energy [31,32]. This result aligns with the findings reported by Du et al., indicating the introduction of strong electronic interactions between Fe, Ni, and Se atoms after Fe doping [31,32].
The electrochemical performance of all materials was characterized in the 1   M KOH electrolyte using the standard three-electrode system, with relevant data presented in Figure 4. The catalysts were uniformly coated onto a 3   m m glassy carbon electrode (GCE), with a total catalyst loading of approximately 0.114   m g / c m 2 . The working electrode potential (vs. Ag/AgCl) was converted to the reversible hydrogen electrode (RHE) scale using the formula: V R H E = 0.197 + 0.059 × p H + V ( A g / A g C l ) . During electrochemical measurements, the working electrode was rotated at 1600   r p m to remove generated oxygen bubbles. Polarization curves for the synthesized samples were recorded at a low scan rate of 5   m V / s to minimize capacitive current effects.
As shown in Figure 4a, all samples exhibited significant catalytic activity for the oxygen evolution reaction (OER). The oxidation peak observed for NiSe2 at approximately 1.35   V before the OER onset potential is attributed to the Ni(OH)2/NiOOH transformation, consistent with literature reports [31,34]. The overpotentials required for the four samples to achieve a current density of 10   m A   c m 2 were statistically analyzed, and the results are shown in Figure 4b. It can be seen that FeSe2 demonstrated the lowest catalytic activity, requiring a high overpotential of 601   m V to achieve a current density of 10   m A   c m 2 . When doped with 0.1   m m o l of Fe2(SO4)3, the resulting Ni0.9Fe0.1Se2 showed significantly enhanced OER performance. However, higher Fe doping levels were not beneficial; increasing the doping amount to 0.2   m m o l resulted in decreased OER performance. Among the four catalyst samples, Ni0.9Fe0.1Se2 exhibited the highest catalytic activity, requiring an overpotential of only 231   m V at a current density of 10   m A   c m 2 . In comparison, the overpotentials for NiSe2 and Ni0.83Fe0.17Se2 were 301   m V and 330   m V , respectively. These results indicate that the Fe/Ni ratio during the reaction not only influences the morphology and composition of the samples but also significantly modulates their OER performance. Notably, Ni0.9Fe0.1Se2 prepared with a low doping amount ( 0.1   m m o l ) exhibited optimal performance, demonstrating superior characteristics compared to various selenides and transition metal compound catalysts reported in the literature [23,33,35,36,37], as shown in Table S1. Furthermore, compared to the higher-cost NiSe2 and the poorly performing FeSe2 in terms of catalytic activity, the doping of a small amount of Fe reduces the cost while enhancing the catalytic activity of the material, resulting in excellent cost-effectiveness (see Table S2). Regarding nickel-based oxygen evolution catalysts, the introduction of iron (Fe)—whether accidental or intentional—has been widely demonstrated to significantly enhance the catalytic activity for the oxygen evolution reaction (OER). The proposed mechanisms for this iron-enhanced performance primarily include the following [27]: First, Fe3+ ions can increase the electrical conductivity of the catalytic film, thereby improving OER performance; second, Fe3+ can acquire sufficient charge transfer capability from conductive substrates (such as nickel or cobalt oxyhydroxides) and conductive supports (such as redox-activated gold), leading to superior electrocatalytic performance; third, Fe3+ sites are regarded as “fast” active centers in OER kinetics due to their near-optimal adsorption energy for OER intermediates; fourth, the incorporation of Fe3+ can suppress metal oxidation steps and promote reduction steps related to oxygen evolution, thereby optimizing the OER process. Nevertheless, a unified and definitive theory regarding the enhancement mechanism of iron has not yet been established. In this study, XPS results indicate that compared to undoped NiSe2, the binding energies of Ni and Se in the iron-doped samples (Ni0.9Fe0.1Se2 and Ni0.83Fe0.17Se2) shift negatively. This observation suggests that the introduction of Fe induces significant reconstruction of the local electronic structure of NiSe2, specifically manifested as electron transfer from Fe to Ni and Se regions, increasing the electron cloud density of Ni and Se atoms. The enhanced catalytic activity of Ni0.9Fe0.1Se2 is likely attributed to this modulation of the electronic structure, which optimizes the adsorption strength of nickel active sites toward key oxygen intermediates (such as *O, *OH, and *OOH) [18,27], bringing it closer to the optimal adsorption energy required for OER. The optimized adsorption energy of intermediates effectively reduces the energy barriers of various steps in the OER pathway, particularly that of the rate-determining step (RDS), thereby significantly lowering the reaction overpotential.
The Tafel slope is a fundamental parameter in electrocatalysis and holds significant importance. Its value directly reflects the ease of the electrode reaction—a lower Tafel slope indicates that a high current density can be achieved at a low overpotential, thereby greatly improving the efficiency of water electrolysis and demonstrating excellent catalytic performance. The OER kinetics were evaluated by converting polarization curves into Tafel plots (potential vs.   l o g ( j ) ). As shown in Figure 5a, the Tafel slope for Ni0.9Fe0.1Se2 is 94   m V   d e c 1 , which is lower than those of Ni0.83Fe0.17Se2 ( 116   m V   d e c 1 ) and FeSe2 ( 565   m V   d e c 1 ), but higher than that of NiSe2 ( 76   m V   d e c 1 ). It can be seen that the Tafel slope of the catalyst exhibits an increasing trend with higher iron doping content. A larger Tafel slope is detrimental to OER kinetics. Although doping trace amounts of iron into nickel selenide improves OER kinetics (as demonstrated by Ni0.9Fe0.1Se2), the observed increase in Tafel slope with higher iron content indicates that the composition of this catalyst still requires further optimization.
Furthermore, the durability of a catalyst is a critical indicator for assessing its potential for practical applications. In this study, cyclic stability tests were conducted on the Ni0.9Fe0.1Se2 sample. As shown in Figure 5b, the results from chronoamperometric testing indicate that after 5 h of continuous operation, the current density decreased from an initial value of 10   m A   c m 2 to 8.3   m A   c m 2 . Although the degree of attenuation is relatively modest, it still suggests that the stability of the material requires further improvement, and the doping strategy needs additional optimization.

3. Materials and Methods

3.1. Chemicals and Materials

Oleylamine (OAm, C18H37N, 80 ~ 90 % ), tetralin (C10H12, > 98 % ), Ni(acac)2 (Ni(C5H7O2)2, > 95 % ), ferric sulfate (Fe2(SO4)3, > 80 % ), ethanol (CH3CH2OH, 99.7 % ), isopropanol (C3H8O, 99 % ), n-hexane (C6H14, 99.7 % ) and n-octylamine (C8H19N, > 99 % ) were purchased from Sinopharm Chemical Reagent Co., Ltd., Shanghai, China. Potassium hydroxide (KOH) was purchased from Aladdin Biochemical Technology Co., Ltd., Shanghai, China. Nafion solution ( 5 % ) was purchased from Sigma-Aldrich® (Shanghai, China). High-purity argon ( > 99.99 % ), high-purity oxygen ( > 99.99 % ) and deionized water ( 18.2   M Ω · c m ) were used in the experiments. All chemical reagents were used without further purification.

3.2. The Synthesis of NiSe2

Dissolve 1   m m o l of Ni(acac)2 in a mixed solution of 10   m L OAm and 10   m L tetralin. Preheat the mixture to 413   K under magnetic stirring and maintain this temperature for 1   h to form the precursor solution. After cooling to room temperature, disperse 1.5   m m o l of selenium powder in a mixture of 2   m L OAm and 2   m L tetralin, then add this dispersion to the precursor solution. Stir for several minutes until a homogeneous black solution forms. Subsequently, heat the solution to 413   K and maintain for 2   h to synthesize NiSe2.

3.3. The Synthesis of FeSe2

FeSe2 was prepared using an analogous procedure to NiSe2 synthesis, replacing Ni(acac)2 with ferric sulfate (Fe2(SO4)3) as the metal precursor.

3.4. The Synthesis of Fe Doped NiSe2

After cooling as-synthesized NiSe2 to room temperature, add a dispersion of Fe2(SO4)3 in 2   m L OAm and 2   m L tetralin to the NiSe2 solution. Continue the reaction at 413   K for 2   h to obtain the Fe-doped NiSe2. By varying the amount of Fe2(SO4)3 added ( 0.1   m m o l , or 0.2   m m o l ), Fex-NiSe2 samples with controlled doping levels were prepared.
All reactions were conducted in a three-necked round-bottomed flask under a continuous argon atmosphere. Finally, all products were washed twice with a mixed solvent of isopropanol and n-hexane, followed by ethanol washing, and dried under vacuum.

3.5. Characteristic Measurement

The phase composition of the sample was analyzed using powder X-ray diffraction (XRD) with the DX-2700 diffractometer. Sample characterization was performed using field emission scanning electron microscopy (FE-SEM, Hitachi S-4800, Hitachi High-Technologies Corporation, Tokyo, Japan). The surface elemental composition and chemical states of the sample were obtained by X-ray photoelectron spectroscopy (XPS) using a Kratos XSAM 800 (Kratos Analytical Ltd., Manchester, UK) system with an Al Kα X-ray source.

3.6. Preparation of the Working Electrode

To remove surface oxides generated during sample storage and enhance its dispersibility, the black powder sample was sonicated in n-octylamine solution ( 2   m g / m L ) for 2 h, followed by standing for 24 h. After washing with deionized water, the product was uniformly dispersed in a mixed solution of isopropanol/water/Nafion (volume ratio 1 : 3 : 0.01 ) to form a catalyst ink. Then, 4   μ L of the ink was drop-cast onto a 3   m m diameter glassy carbon electrode (GCE) polished with Al2O3 powder, yielding a catalyst loading of approximately 0.114   m g / c m 2 . After natural evaporation of deionized water and isopropanol and thorough drying of the sample on the electrode, a uniform catalyst layer was formed.

3.7. Electrochemical Measurements

All electrochemical measurements were performed at room temperature using a CHI 760e electrochemical workstation (CH Instruments, Inc., Shanghai, China) with a standard three-electrode configuration. Using an Ag/AgCl electrode (in 3   M KCl) as the reference electrode and a platinum wire as the counter electrode, the electrolyte was a 1   M KOH solution. Prior to testing, the sample was activated with 30 cycles of CV scanning at a scan rate of 50   m V / s to ensure a thorough evaluation of its electrochemical performance. During testing, oxygen was continuously bubbled into the electrolyte. Cyclic voltammetry (CV), linear sweep voltammetry (LSV), and Tafel measurements were performed at a rotation rate of 1600   r p m with a scan rate of 5   m V / s . The electrocatalytic stability of the samples was evaluated by the chronoamperometric (CA) stability test at fixed potential of 1.46   V vs. RHE.

4. Conclusions

This study compared the performance of four catalysts: NiSe2, FeSe2, and Ni0.9Fe0.1Se2 and Ni0.83Fe0.17Se2 synthesized with two different Fe2(SO4)3 doping levels. The OER catalytic activity was found to follow the order: N i 0.9 F e 0.1 S e 2 > N i S e 2 > N i 0.83 F e 0.17 S e 2 > F e S e 2 . These results indicate that low-level iron doping is an effective strategy to enhance the OER catalytic activity of NiSe2. Morphological and compositional characterization of the samples before and after doping revealed that the iron-doped catalysts developed a uniform coating on the originally nano-dendritic structure. Moreover, XPS results showed negative shifts in the binding energies of both Ni and Se, which can be attributed to a significant reconstruction of the local electronic structure of NiSe2 due to the incorporation of Fe. This reconstruction induces strong electronic interactions among Fe, Ni, and Se, resulting in a notable synergistic effect. As a result, the optimized catalyst requires an overpotential of only 231   m V to achieve a current density of 10   m A   c m 2 . These findings demonstrate that iron doping is an effective approach for improving the OER performance of nickel selenide. It is worth noting that the optimal iron doping level in this study ( ~ 10   a t % ) is significantly lower than that reported by Du et al. for Ni1.12Fe0.49Se2 ( ~ 49   a t % ). This study provides a broader research foundation for developing high-performance electrocatalysts based on transition metal selenides (and sulfides) through doping strategies. Furthermore, a wider range of iron doping levels should be considered in subsequent research.

Supplementary Materials

The following supporting information can be downloaded at https://www.mdpi.com/article/10.3390/catal15090876/s1. Figure S1: EDS spectra of samples obtained with different addition amounts of Fe2(SO4)3 ( 0.1   m m o l vs 0.2   m m o l ): (a) 0.1   m m o l Fe2(SO4)3, with an elemental composition of S ( 1.8   a t % ), O ( 10.2   a t % ), Fe ( 2.6   a t % ), Ni ( 23.4   a t % ), and Se ( 62   a t % ); (b) 0.2   m m o l Fe2(SO4)3, with an elemental composition of S ( 2.1   a t % ), O ( 11.8   a t % ), Fe ( 4.4   a t % ), Ni ( 21.7   a t % ), and Se ( 60   a t % ); Figure S2: SEM-EDS mapping images reflects Fe, Ni and Se atom distributions for the Ni0.9Fe0.1Se2; Table S1: Comparison of the electrocatalytic OER performance of Ni0.9Fe0.1Se2 with other reported transition metal compounds and Fe-doped selenides; Table S2: Comparison of cost and application potential of three nanoelectrocatalytic materials. References [38,39] are cited in the supplementary materials.

Author Contributions

Conceptualization, L.S.; methodology, Y.M.; software, Y.M.; validation, L.S., Y.M. and B.L.; formal analysis, L.S. and Y.M.; investigation, L.S.; resources, Y.M.; data curation, Y.M.; writing—original draft preparation, L.S.; writing—review and editing, L.S. and Y.M.; visualization, L.S.; supervision, L.S., Y.M. and B.L.; project administration, B.L.; funding acquisition, B.L. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Natural Science Foundation of China (NSFC, Grant No. 12105037).

Data Availability Statement

The original contributions presented in this study are included in the article. Further inquiries can be directed to the corresponding author(s).

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Jamesh, M.I. Recent progress on earth abundant hydrogen evolution reaction and oxygen evolution reaction bifunctional electrocatalyst for overall water splitting in alkaline media. J. Power Sources 2016, 333, 213–236. [Google Scholar] [CrossRef]
  2. Turner, J.A. Sustainable hydrogen production. Science 2004, 305, 972–974. [Google Scholar] [CrossRef]
  3. Cabán-Acevedo, M.; Stone, M.L.; Schmidt, J.; Thomas, J.G.; Ding, Q.; Chang, H.-C.; Tsai, M.-L.; He, J.-H.; Jin, S. Efficient hydrogen evolution catalysis using ternary pyrite-type cobalt phosphosulphide. Nat. Mater. 2015, 14, 1245–1251. [Google Scholar] [CrossRef] [PubMed]
  4. Bajdich, M.; García-Mota, M.; Vojvodic, A.; Nørskov, J.K.; Bell, A.T. Theoretical investigation of the activity of cobalt oxides for the electrochemical oxidation of water. J. Am. Chem. Soc. 2013, 135, 13521–13530. [Google Scholar] [CrossRef]
  5. Liu, T.; Lu, J.; Chen, Z.; Luo, Z.; Ren, Y.; Zhuge, X.; Luo, K.; Ren, G.; Lei, W.; Liu, D. Advances, mechanisms and applications in oxygen evolution electrocatalysis of gold-driven. Chem. Eng. J. 2024, 496, 153719. [Google Scholar] [CrossRef]
  6. Si, C.; Zhang, W.; Lu, Q.; Guo, E.; Yang, Z.; Chen, J.; He, X.; Luo, J. Recent advances in perovskite catalysts for efficient overall water splitting. Catalysts 2022, 12, 601. [Google Scholar] [CrossRef]
  7. Wu, L.; Yu, L.; Zhang, F.; McElhenny, B.; Luo, D.; Karim, A.; Chen, S.; Ren, Z. Heterogeneous bimetallic phosphide Ni2P-Fe2P as an efficient bifunctional catalyst for water/seawater splitting. Adv. Funct. Mater. 2021, 31, 2006484. [Google Scholar] [CrossRef]
  8. Meena, A.; Thangavel, P.; Jeong, D.S.; Singh, A.N.; Jana, A.; Im, H.; Nguyen, D.A.; Kim, K.S. Crystalline-amorphous interface of mesoporous Ni2P@FePOxHy for oxygen evolution at high current density in alkaline-anion-exchange-membrane water-electrolyzer. Appl. Catal. B Environ. 2022, 306, 121127. [Google Scholar] [CrossRef]
  9. Guo, K.; Wang, Y.; Yang, S.; Huang, J.; Zou, Z.; Pan, H.; Shinde, P.S.; Pan, S.; Huang, J.; Xu, C. Bonding interface boosts the intrinsic activity and durability of NiSe@Fe2O3 heterogeneous electrocatalyst for water oxidation. Sci. Bull. 2021, 66, 52–61. [Google Scholar] [CrossRef]
  10. Kwak, I.H.; Im, H.S.; Jang, D.M.; Kim, Y.W.; Park, K.; Lim, Y.R.; Cha, E.H.; Park, J. CoSe2 and NiSe2 nanocrystals as superior bifunctional catalysts for electrochemical and photoelectrochemical water splitting. ACS Appl. Mater. Interfaces 2016, 8, 5327–5334. [Google Scholar] [CrossRef] [PubMed]
  11. Enez, S.; Karani Konuksever, V.; Samuei, S.; Karadas, F.; Ülker, E. Enhancing oxygen evolution catalytic performance of nickel borate with cobalt doping and carbon nanotubes. ChemistrySelect 2023, 8, e202203561. [Google Scholar] [CrossRef]
  12. Gupta, S.; Patel, M.K.; Miotello, A.; Patel, N. Metal boride-based catalysts for electrochemical water-splitting: A review. Adv. Funct. Mater. 2020, 30, 1906481. [Google Scholar] [CrossRef]
  13. Ren, J.-T.; Chen, L.; Wang, H.-Y.; Tian, W.-W.; Song, X.-L.; Kong, Q.-H.; Yuan, Z.-Y. Synergistic Activation of Crystalline Ni2P and Amorphous NiMoOx for Efficient Water Splitting at High Current Densities. ACS Catal. 2023, 13, 9792–9805. [Google Scholar] [CrossRef]
  14. Li, Y.M.; He, H.Y.; Fu, W.; Mu, C.Z.; Tang, X.Z.; Liu, Z.; Chi, D.Z.; Hu, X. In-grown structure of NiFe mixed metal oxides and CNT hybrid catalysts for oxygen evolution reaction. Chem. Commun. 2016, 52, 1439–1442. [Google Scholar] [CrossRef]
  15. Thangavel, P.; Kim, G.; Kim, K.S. Electrochemical integration of amorphous NiFe (oxy) hydroxides on surface-activated carbon fibers for high-efficiency oxygen evolution in alkaline anion exchange membrane water electrolysis. J. Mater. Chem. A 2021, 9, 14043–14051. [Google Scholar] [CrossRef]
  16. Jiang, M.; Zhai, H.; Chen, L.; Mei, L.; Tan, P.; Yang, K.; Pan, J. Unraveling the synergistic mechanism of bi-functional nickel–iron phosphides catalysts for overall water splitting. Adv. Funct. Mater. 2023, 33, 2302621. [Google Scholar] [CrossRef]
  17. Xia, X.; Wang, L.; Sui, N.; Colvin, V.L.; Yu, W.W. Recent progress in transition metal selenide electrocatalysts for water splitting. Nanoscale 2020, 12, 12249–12262. [Google Scholar] [CrossRef] [PubMed]
  18. Deng, R.; Yao, H.; Wang, Y.; Wang, C.; Zhang, S.; Guo, S.; Li, Y.; Ma, S. Interface effect of Fe doped NiSe/Ni3Se2 heterojunction as highly efficient electrocatalysts for overall water splitting. Chem. Eng. J. 2024, 488, 150996. [Google Scholar] [CrossRef]
  19. Dang, N.K.; Tiwari, J.N.; Sultan, S.; Meena, A.; Kim, K.S. Multi-site catalyst derived from Cr atoms-substituted CoFe nanoparticles for high-performance oxygen evolution activity. Chem. Eng. J. 2021, 404, 126513. [Google Scholar] [CrossRef]
  20. Sultan, S.; Ha, M.; Kim, D.Y.; Tiwari, J.N.; Myung, C.W.; Meena, A.; Shin, T.J.; Chae, K.H.; Kim, K.S. Superb water splitting activity of the electrocatalyst Fe3Co(PO4)4 designed with computation aid. Nat. Commun. 2019, 10, 5195. [Google Scholar] [CrossRef]
  21. Li, S.; Gao, Y.; Li, N.; Ge, L.; Bu, X.; Feng, P. Transition metal-based bimetallic MOFs and MOF-derived catalysts for electrochemical oxygen evolution reaction. Energy Environ. Sci. 2021, 14, 1897–1927. [Google Scholar] [CrossRef]
  22. Gao, R.; Zhang, H.; Yan, D. Iron diselenide nanoplatelets: Stable and efficient water-electrolysis catalysts. Nano Energy 2017, 31, 90–95. [Google Scholar] [CrossRef]
  23. Swesi, A.T.; Masud, J.; Nath, M. Nickel selenide as a high-efficiency catalyst for oxygen evolution reaction. Energy Environ. Sci. 2016, 9, 1771–1782. [Google Scholar] [CrossRef]
  24. Cai, C.; Mi, Y.; Han, S.; Wang, Q.; Liu, W.; Wu, X.; Zheng, Z.; Xia, X.; Qiao, L.; Zhou, W.; et al. Engineering ordered dendrite-like nickel selenide as electrocatalyst. Electrochim. Acta 2019, 295, 92–98. [Google Scholar] [CrossRef]
  25. Zhou, H.; Wang, Y.; He, R.; Yu, F.; Sun, J.; Wang, F.; Lan, Y.; Ren, Z.; Chen, S. One-step synthesis of self-supported porous NiSe2/Ni hybrid foam: An efficient 3D electrode for hydrogen evolution reaction. Nano Energy 2016, 20, 29–36. [Google Scholar] [CrossRef]
  26. Tichenor, R.L. Nickel oxides-relation between electrochemical and foreign ion content. Ind. Eng. Chem. 1952, 44, 973–977. [Google Scholar] [CrossRef]
  27. Anantharaj, S.; Kundu, S.; Noda, S. “The Fe Effect”: A review unveiling the critical roles of Fe in enhancing OER activity of Ni and Co based catalysts. Nano Energy 2021, 80, 105514. [Google Scholar] [CrossRef]
  28. Trotochaud, L.; Ranney, J.K.; Williams, K.N.; Boettcher, S.W. Solution-cast metal oxide thin film electrocatalysts for oxygen evolution. J. Am. Chem. Soc. 2012, 134, 17253–17261. [Google Scholar] [CrossRef]
  29. Trotochaud, L.; Young, S.L.; Ranney, J.K.; Boettcher, S.W. Nickel–Iron Oxyhydroxide Oxygen-Evolution Electrocatalysts: The Role of Intentional and Incidental Iron Incorporation. J. Am. Chem. Soc. 2014, 136, 6744–6753. [Google Scholar] [CrossRef]
  30. Burke, M.S.; Kast, M.G.; Trotochaud, L.; Smith, A.M.; Boettcher, S.W. Cobalt–Iron (Oxy)hydroxide Oxygen Evolution Electrocatalysts: The Role of Structure and Composition on Activity, Stability, and Mechanism. J. Am. Chem. Soc. 2015, 137, 3638–3648. [Google Scholar] [CrossRef]
  31. Du, Y.; Cheng, G.; Luo, W. Colloidal synthesis of urchin-like Fe doped NiSe2 for efficient oxygen evolution. Nanoscale 2017, 9, 6821–6825. [Google Scholar] [CrossRef] [PubMed]
  32. Du, Y.; Cheng, G.; Luo, W. NiSe2/FeSe2 nanodendrites: A highly efficient electrocatalyst for oxygen evolution reaction. Catal. Sci. Technol. 2017, 7, 4604–4608. [Google Scholar] [CrossRef]
  33. Chang, K.; Tran, D.T.; Wang, J.; Kim, N.H.; Lee, J.H. A 3D hierarchical network derived from 2D Fe-doped NiSe nanosheets/carbon nanotubes with enhanced OER performance for overall water splitting. J. Mater. Chem. A 2022, 10, 3102–3111. [Google Scholar] [CrossRef]
  34. Louie, M.W.; Bell, A.T. An investigation of thin-film Ni–Fe oxide catalysts for the electrochemical evolution of oxygen. J. Am. Chem. Soc. 2013, 135, 12329–12337. [Google Scholar] [CrossRef]
  35. Chen, S.; Kang, Z.; Hu, X.; Zhang, X.; Wang, H.; Xie, J.; Zheng, X.; Yan, W.; Pan, B.; Xie, Y. Delocalized spin states in 2D atomic layers realizing enhanced electrocatalytic oxygen evolution. Adv. Mater. 2017, 29, 1701687. [Google Scholar] [CrossRef]
  36. Yu, X.-Y.; Feng, Y.; Guan, B.; Lou, X.W.D.; Paik, U. Carbon coated porous nickel phosphides nanoplates for highly efficient oxygen evolution reaction. Energy Environ. Sci. 2016, 9, 1246–1250. [Google Scholar] [CrossRef]
  37. Gao, M.; Sheng, W.; Zhuang, Z.; Fang, Q.; Gu, S.; Jiang, J.; Yan, Y. Efficient water oxidation using nanostructured α-nickel-hydroxide as an electrocatalyst. J. Am. Chem. Soc. 2014, 136, 7077–7084. [Google Scholar] [CrossRef] [PubMed]
  38. Stern, L.-A.; Feng, L.; Song, F.; Hu, X. Ni2P as a Janus catalyst for water splitting: The oxygen evolution activity of Ni2P nanoparticles. Energy Environ. Sci. 2015, 8, 2347–2351. [Google Scholar] [CrossRef]
  39. Li, J.; Li, J.; Zhou, X.; Xia, Z.; Gao, W.; Ma, Y.; Qu, Y. Highly efficient and robust nickel phosphides as bifunctional electrocatalysts for overall water-splitting. ACS Appl. Mater. Interfaces 2016, 8, 10826–10834. [Google Scholar] [CrossRef]
Figure 1. XRD patterns of samples prepared with different solutes: (a) Ni(acac)2 and selenium powder, (b) Fe2(SO4)3 and selenium powder, (c) Ni(acac)2, selenium powder, and Fe2(SO4)3.
Figure 1. XRD patterns of samples prepared with different solutes: (a) Ni(acac)2 and selenium powder, (b) Fe2(SO4)3 and selenium powder, (c) Ni(acac)2, selenium powder, and Fe2(SO4)3.
Catalysts 15 00876 g001
Figure 2. SEM images of (a) NiSe2, (b) FeSe2, (c) Ni0.9Fe0.1Se2, and (d) Ni0.83Fe0.17Se2.
Figure 2. SEM images of (a) NiSe2, (b) FeSe2, (c) Ni0.9Fe0.1Se2, and (d) Ni0.83Fe0.17Se2.
Catalysts 15 00876 g002
Figure 3. (a) The full XPS spectra of Ni0.9Fe0.1Se2. (bd) The high-resolution Fe2p, Ni2p and Se3d XPS spectra of Ni0.9Fe0.1Se2. (e,f) The high-resolution Se3d and Ni2p XPS spectra of Ni0.9Fe0.1Se2, Ni0.83Fe0.17Se2, FeSe2, and NiSe2 samples.
Figure 3. (a) The full XPS spectra of Ni0.9Fe0.1Se2. (bd) The high-resolution Fe2p, Ni2p and Se3d XPS spectra of Ni0.9Fe0.1Se2. (e,f) The high-resolution Se3d and Ni2p XPS spectra of Ni0.9Fe0.1Se2, Ni0.83Fe0.17Se2, FeSe2, and NiSe2 samples.
Catalysts 15 00876 g003
Figure 4. (a) Polarization curves of NiSe2, FeSe2, Ni0.9Fe0.1Se2, and Ni0.83Fe0.17Se2. (b) The overpotentials required to achieve a current density of 10   m A   c m 2 for the samples NiSe2, FeSe2, Ni0.9Fe0.1Se2, and Ni0.83Fe0.17Se2.
Figure 4. (a) Polarization curves of NiSe2, FeSe2, Ni0.9Fe0.1Se2, and Ni0.83Fe0.17Se2. (b) The overpotentials required to achieve a current density of 10   m A   c m 2 for the samples NiSe2, FeSe2, Ni0.9Fe0.1Se2, and Ni0.83Fe0.17Se2.
Catalysts 15 00876 g004
Figure 5. (a) Tafel plots of NiSe2, FeSe2, Ni0.9Fe0.1Se2, and Ni0.83Fe0.17Se2. (b) Variation in current density with time for the Ni0.9Fe0.1Se2 under constant potential.
Figure 5. (a) Tafel plots of NiSe2, FeSe2, Ni0.9Fe0.1Se2, and Ni0.83Fe0.17Se2. (b) Variation in current density with time for the Ni0.9Fe0.1Se2 under constant potential.
Catalysts 15 00876 g005
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

Sun, L.; Mi, Y.; Li, B. Iron-Doped NiSe2 and Its Enhanced Oxygen Evolution Reaction Activity. Catalysts 2025, 15, 876. https://doi.org/10.3390/catal15090876

AMA Style

Sun L, Mi Y, Li B. Iron-Doped NiSe2 and Its Enhanced Oxygen Evolution Reaction Activity. Catalysts. 2025; 15(9):876. https://doi.org/10.3390/catal15090876

Chicago/Turabian Style

Sun, Lijie, Yaqun Mi, and Bo Li. 2025. "Iron-Doped NiSe2 and Its Enhanced Oxygen Evolution Reaction Activity" Catalysts 15, no. 9: 876. https://doi.org/10.3390/catal15090876

APA Style

Sun, L., Mi, Y., & Li, B. (2025). Iron-Doped NiSe2 and Its Enhanced Oxygen Evolution Reaction Activity. Catalysts, 15(9), 876. https://doi.org/10.3390/catal15090876

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