Jianliang
Yuan
ad,
Qianglong
Qi
ab,
Qingwen
Wan
ab,
Jiangli
Gong
ad,
Yue
Zhang
ab,
Yuebin
Feng
c,
Chengxu
Zhang
*a and
Jue
Hu
*ab
aFaculty of Metallurgical and Energy Engineering, Kunming University of Science and Technology, Kunming, China. E-mail: chxzhang@kust.edu.cn; hujue@kust.edu.cn
bKey Laboratory of Unconventional Metallurgy, Kunming University of Science and Technology, Kunming, Yunnan, China
cFaculty of Science, Kunming University of Science and Technology, Kunming, China
dLuXi KuoBo Precious Metals Co. Ltd., Honghe, Yunnan, China
First published on 25th February 2025
Medium-entropy alloys (MEAs) as electrocatalysts have attracted considerable attention in the field of water splitting. However, effective modulation of MEAs to achieve highly efficient catalysis remains a challenge. Herein, we applied a metal–organic framework (MOF) templating strategy to obtain FeCoNi MEA nanoparticles with excellent oxygen evolution reaction (OER) activity and tungsten improved FeCoNi-W medium-entropy heterostructure catalysts. The incorporation of tungsten changes the electronic structure of FeCoNi MEA. The mesoporous alloy exhibits multiple active sites and unique atomic-level synergies that enhance the effective binding of reactants and the formation of crucial *OH intermediates critical for OER. The rationally designed and constructed tungsten-refined FeCoNi-W medium-entropy heterostructure electrode demonstrates superior OER performance (270 mV at 10 mA cm−2, a Tafel slope of 43.2 mV dec−1) and stability (50 h at 100 mA cm−2) compared to commercial noble metal electrodes. This work will provide a basis for tailoring the properties of medium/high-entropy alloys (M/HEAs) through local chemical modification.
Here, we synthesized a nano-heterojunction catalyst rich in rare earth elements Fe, Co, Ni, and W using a simple hydrothermal method. By precisely controlling the equimolar ratio of transition metal elements, we achieved high electrocatalytic performance for FeCoNi-W. The prepared electrocatalysts exhibited excellent catalytic performance, benefiting from abundant active sites, multi-element synergistic effects, and entropy stability. In particular, we rationally designed and constructed a tungsten-refined FeCoNi-W medium-entropy heterostructure electrode that surpassed the FeCoNi MEA electrode in terms of OER performance and stability (achieving an overpotential of 270 mV at 10 mA cm−2 with a high turnover frequency and stable operation for 50 hours at 100 mA cm−2). This superior performance is attributed to the addition of tungsten, which alloyed FeCoNi with the less electronegative W element, forming a unique medium-entropy heterostructure. This process fine-tunes the surface electronic states of the active metal centers, facilitating the adsorption of OH and enhancing the effective binding of reactants. This study provides a simple and effective method for synthesizing heterogeneous materials containing nanoscale multi-metal elements, providing new vitality into the development of the hydrogen energy field.
![]() | ||
Fig. 1 (a) XRD patterns and (b) Raman spectra of FeCoNi MEA and FeCoNi-W medium-entropy heterostructures. |
A detailed investigation of the microstructure and morphology of the catalysts was conducted using transmission electron microscopy (TEM) and scanning electron microscopy (SEM). The FeCoNi and FeCoNi-W medium-entropy heterostructure samples exhibit similar morphological characteristics, consisting of aggregated particles formed by the stacking of irregularly surfaced nanoparticles (Fig. 2a, b and Fig. S1†). To gain a deeper understanding of the microstructure of these samples, High-Resolution TEM (HR-TEM) was employed. The HR-TEM images of FeCoNi-W presented in Fig. 2c and d clearly reveal the abundant presence of FeWO4 nanoparticles on the surfaces and edges of the FeCoNi nanospheres. In Fig. 2e, lattice fringes with spacings of approximately 2.94 Å (yellow lines) and 2.06 Å (red lines) were observed, which are attributed to the (111) plane of medium-entropy oxides within the medium-entropy heterostructure nanocatalyst and the (111) diffraction peak of the FeCoNi alloy nanocatalyst, respectively. Notably, the combination of FeWO4 and FeCoNi forms a unique heterogeneous structure, resulting in a shift in the peak position of FeWO4 (111) compared to pure FeCoNi. To gain insights into the elemental distribution of FeCoNi-W, energy-dispersive X-ray (EDX) mapping analysis was performed. The EDX Spectroscopy elemental mapping results of the FeCoNi-W sample show a uniform distribution of five elements within the sample, further confirming the successful synthesis of the FeCoNi-W heterogeneous structure (Fig. 2f and Fig. S2†).
X-ray photoelectron spectroscopy (XPS) characterization and analysis were performed to investigate in detail the surface chemical states and chemical structures of the prepared FeCoNi MEA and the alloyed FeCoNi-W MEA/MEO catalyst with a heterogeneous compound formed by incorporating the less electronegative element W. As shown in Fig. S3,† the XPS spectra of both FeCoNi MEA and FeCoNi-W MEA/MEO show the presence of the same constituents, with a distinct peak for element W in FeCoNi-W MEA/MEO clearly visible. In addition, peak matching was performed on the high-resolution XPS spectra of the primary constituents of the FeCoNi MEA and FeCoNi-W MEA/MEO catalysts. The Fe 2p high-resolution spectrum (Fig. 3a) exhibits two peaks at 710.8 eV and 724.91 eV, corresponding to Fe2+ 2p3/2 and Fe2+ 2p1/2, respectively, while peaks at 713.6 eV and 727.48 eV are associated with Fe3+ in the catalyst material. The peaks located at 709.08 eV and 723.23 eV represent metallic Fe, and those at 716.7 eV and 734.07 eV are satellite peaks.28 Furthermore, surface oxidation of alloy particles leads to the formation of high-valent iron. Compared with the FeCoNi MEA sample, the binding energy of the Fe 2p3/2 peak in the FeCoNi-W MEA/MEO catalyst exhibits a positive shift (to 707.3 eV), indicating that the incorporation of the less electronegative W element reduces the electron cloud density around Fe atoms, effectively “pulling away” or “depleting” electrons, thereby causing a positive shift in the binding energy of the Fe 2p3/2 peak. This alteration in the electronic structure may influence the catalytic performance of the FeCoNi-W catalyst. As a heterophase element, W possesses a different atomic radius and electron configuration compared to Fe, Co, and Ni, disrupting the original electronic balance and causing a redistribution of electrons around Fe atoms when the W heterophase is formed in the alloy. Analysis of the high-resolution Ni XPS spectrum reveals two prominent peaks in the Ni 2p3/2 region of the FeCoNi-W catalyst: the peak at 853.2 eV corresponds to metallic Ni0,29 while the peak at 856.9 eV is attributed to Ni2+, indicating partial oxidation on the catalyst surface. Notably, the introduction of the less electronegative W element to form an alloy/oxide heterogeneous junction results in a slight negative shift in the binding energy of metallic Ni. Additionally, in the spectrum of FeCoNi, the peaks at 781.0 eV and 796.7 eV correspond to the 2p3/2 and 2p1/2 energy levels of metallic Co, respectively (Fig. 3b).30,31 Analysis of the high-resolution Co XPS spectrum reveals two prominent peaks in the Co 2p3/2 region of the FeCoNi-W catalyst: the peak at 778.8 eV corresponds to metallic Co0, while the peak at 781.1 eV is attributed to Co2+, indicating partial oxidation on the catalyst surface. For the FeCoNi-W catalyst, the peaks at 853.2 eV and 870.4 eV belong to the 2p3/2 and 2p1/2 energy levels of metallic Ni0, respectively (Fig. 3c). It is worth noting that the incorporation of the less electronegative W element into the FeCoNi alloy results in varying degrees of binding energy shifts for Fe0, Co0, and Ni0. This phenomenon reveals a significant synergistic electronic coupling effect among Fe, Co, Ni, and W atoms (Fig. 3d), which may alter the adsorption behavior of OER intermediates on Co, Ni, and Fe sites.32 This discovery further confirms the existence of synergistic electronic coupling among Fe, Co, Ni, and W atoms.
The OER catalytic activity was investigated in a 1.0 M KOH electrolyte employing a standard three-electrode setup. Subsequently, the electrocatalytic OER activities of FeCoNi MEA and FeCoNi-W MEA/MEO heterostructure nanocatalysts were then compared using cyclic voltammetry (CV) at a scan rate of 0.005 V s−1. The results revealed that the FeCoNi-W heterostructure nanocatalyst exhibited superior performance (Fig. 4a). Notably, the constructed medium-entropy heterogeneous structured electrode of tungsten-refined FeCoNi-W exhibits superior OER performance. At a current density of 10 mA cm−2, the alloying of the less electronegative W element results in the formation of a unique FeCoNi-W medium-entropy alloy/oxide heterogeneous junction catalyst, exhibiting an overpotential of only 270 mV. This is 18.2 mV lower than the overpotential of the FeCoNi MEA. By analyzing the CV curves in the double-layer region at various scan rates (Fig. 4b), the double-layer capacitance (Cdl), which is positively correlated with ESCA, was determined. The heterojunction catalyst obtained after incorporating the less electronegative W element exhibited a higher Cdl value (2.14 mF cm−2 for FeCoNi-W vs. 1.23 mF cm−2 for FeCoNi) compared to the pure FeCoNi MEA. Moreover, FeCoNi-W possessed the lowest Tafel slope (43.2 mV dec−1) (Fig. 4c), indicating that the heterojunction catalyst, with the incorporation of the less electronegative W element, demonstrated faster kinetics during the OER. Furthermore, electrochemical impedance spectroscopy (EIS) was employed to gain insights into the interfacial transport mechanisms of the electrocatalyst. In contrast, the FeCoNi-W MEA/MEO heterostructure nanocatalyst exhibits a lower charge transfer resistance, indicating a superior charge transfer rate during the OER process (Fig. 4d).
As shown in Fig. 5a, b and Fig. S4, S5,† the methanol oxidation reaction (MOR) is employed as an effective method to evaluate the adsorption of OER intermediates, where the increase in current density positively correlates with the coverage of the catalytic reaction intermediate OH*. The results indicate that the FeCoNi-W MEA/MEO heterostructure nanocatalyst exhibits a notably higher current density compared to FeCoNi MEA, suggesting that the formation of heterogeneous interfaces effectively modulates the adsorption of the intermediate OH* in the FeCoNi-W MEA/MEO heterostructure nanocatalyst.33 Additionally, the turnover frequency (TOF) of the FeCoNi MEA, FeCoNi-W MEA/MEO heterostructure nanocatalyst was evaluated to assess its intrinsic activity at a constant overpotential of 300 mV (Fig. 5c and d). The TOF values for Fe and Co in the FeCoNi catalyst are 0.041 s−1 and 0.071 s−1, respectively. However, in the FeCoNi-W catalyst, the TOF values for Fe and Co are significantly higher, at 0.144 s−1 and 0.235 s−1, respectively, representing a 2–3 times increase in intrinsic activity compared to the FeCoNi catalyst. Electrocatalysts, serving as pivotal materials in electrochemical reactions, have their performance directly influencing the efficiency and sustainability of these reactions. During extended periods of use, particularly under high current density operating conditions, the durability of electrocatalysts emerges as one of the crucial indicators for assessing their performance quality. To further investigate the stability of FeCoNi MEA and FeCoNi-W MEA/MEO heterostructure nanocatalysts, we conducted chronopotentiometry experiments. The long-term stability of the FeCoNi-W MEA/MEO heterostructure nanocatalyst material was evaluated through chronopotentiometry (E–t) tests conducted over 50 hours at a constant current density of 100 mA cm−2. As shown in Fig. 6a, after 50 hours of stability testing, the FeCoNi-W material exhibited excellent stability at 10 mA cm−2, with only a slight increase in overpotential of a few millivolts at a current density of 100 mA cm−2, a trend that is significantly superior to that of IrO2 (Fig. 6b and c). Notably, the performance of the FeCoNi-W electrode surpasses many previously reported OER catalysts, such as polymetallic alloys, polymetallic (oxygen) hydroxides,34,35 oxides,36 sulfides,37 and noble metal alloys (Fig. 6d and Tables S1 and S2†).38
Footnote |
† Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d5dt00258c |
This journal is © The Royal Society of Chemistry 2025 |