Sumei
Han‡
*,
Lei
Xu‡
,
Chaoqun
Ma‡
,
Wenbin
Cao
* and
Qipeng
Lu
*
School of Materials Science and Engineering, University of Science and Technology Beijing, Beijing 100083, China. E-mail: sumei@ustb.edu.cn
First published on 4th August 2025
High-entropy alloys (HEAs) have been used as critical electrocatalysts. However, achieving precise atomic-level control over their dimensions and morphologies remains a formidable challenge. In this work, distinctive PtPdRuFeCoNi HEA nanowires (NWs) that possess an average diameter of 1.36 ± 0.05 nm are successfully synthesized. The prepared HEA NWs can serve as a bi-functional electrocatalyst for the hydrogen evolution reaction (HER) and methanol oxidation reaction (MOR) in alkaline solution. The HEA NWs exhibit an ultrasmall overpotential of 14 mV at 10 mA cm−2 for alkaline HER and long-term durability over 50 h. Furthermore, the HEA NWs exhibit excellent catalytic performance for the MOR with a mass activity of 10.4 A mgPGM−1, 13.0 times as high as that of commercial Pt/C. This work offers a facile approach for the synthesis of HEA NWs and facilitates the use of highly active HEA-based catalysts in clean energy conversion and utilization.
The unique composition and intricate surface structure of high-entropy alloys (HEAs) endow them with outstanding physical and chemical properties, making them highly appealing as potential catalysts.19–22 The incorporation of diverse elements facilitates the formation of a solid solution phase, effectively preventing phase segregation. Moreover, multiple elements can create synergistic effects and further lead to some unique catalytic properties of HEAs.23–26 Compared to traditional catalysts, HEAs exhibit superior performance and stability in a wide range of energy conversion reactions such as the MOR, HER, hydrogen oxidation reaction (HOR), oxygen reduction reaction, oxygen evolution reaction, nitrogen reduction reaction and CO2 reduction reaction. However, challenges persist in HEA catalyst development. Synthesis methods are often complex or non-generalizable, and the structural information and catalytic mechanisms of HEAs remain insufficiently understood.27,28 For catalytic reactions, ultrathin nanowires (generally <2 nm) with large surface areas can expose more active sites to achieve high utilization of active atoms.25,29 However, the precise tailoring of dimensions and morphology still faces great challenges, which severely impedes their applications.8,10,25,30 Hence, developing simple and efficient methods to create HEAs with unique morphologies that can effectively improve the catalytic performance is of great importance.
Herein, we synthesize PtPdRuFeCoNi HEA nanowires (NWs) with an ultra-small average diameter of 1.36 ± 0.05 nm through a facile co-reduction method, and the obtained electrocatalysts show high activity for alkaline HER and MOR. The HEA NWs exhibit a remarkably low overpotential for the HER (i.e. 14 mV at 10 mA cm−2 in 1.0 M KOH). We further investigate the electrocatalytic activity of HEA NWs for the MOR, revealing a mass activity of 10.4 A mgPGM−1. This value is 13.0 times the mass activity of commercial Pt/C. Additionally, the mass activity for the MOR remains virtually unchanged even after 20000 s of continuous i–t testing, underscoring the exceptional stability of HEA NWs. The excellent performance of the HEA NWs results from their good intrinsic activity and abundant active sites.
The high-resolution XPS results confirm the coexistence of Pt, Pd, Ru, Ni, Fe, and Co elements, providing additional evidence for the formation of the HEA NWs (Fig. S5†). In Fig. S5a,† two distinct peaks at 71.4 and 74.7 eV are observed in the XPS spectrum of Pt 4f, confirming the existence of metallic Pt within the HEA NWs. Moreover, the two small peaks at 72.3 and 75.6 eV correspond to Pt2+. The above results demonstrate that the Pt in the HEA NWs mainly exists in the metallic state. Similarly, the majority of Pd and Ru are in the metallic state as well (Fig. S5b and c†). As for Fe, Co and Ni, a large part of them are in the oxidation state due to the relatively weak antioxidant ability of non-noble metals (Fig. S5d–f†). It should be noted that satellite peaks exist in the XPS spectra of Co 2p and Ni 2p.32 To investigate the composition of HEA NWs, inductively coupled plasma optical emission spectroscopy (ICP-OES) measurements were conducted. As shown in Table S2,† the ratio of different atoms in HEA NWs is Pt/Ru/Pd/Co/Fe/Ni = 20.8/7.8/13.9/23.4/15.2/19.0, which aligns closely with the data from SEM-EDS analysis.
Notably, W(CO)6 is essential for the preparation of HEA NWs. To confirm this, products are synthesized under the same conditions except for the use of W(CO)6. As presented in Fig. S6,† the shape of the products resembles nanoparticles (NPs) and the average diameter of the NPs is about 10.5 ± 0.2 nm. And the lattice spacing of the (111) plane in the NPs (2.24 Å) is the same as that in HEA NWs (Fig. S7†). Moreover, the XRD pattern of the NPs is mostly the same as that of HEA NWs (Fig. S8†) without the characteristic peaks of pure Pt, Ru, Pd, Co, Fe and Ni. Based on the ICP-OES result, the element composition in NPs is similar to that of HEA NWs with an atomic ratio of Pt/Ru/Pd/Co/Fe/Ni = 19.7/5.9/12.7/22.7/13.4/21.5 (Fig. S9†). The high-resolution XPS spectra indicate that the noble metals are mainly in the metal states and a large part of the oxidized state exists in the non-noble metals (Fig. S10†). Furthermore, W(CO)6 strongly affects the uniformity of the products. As shown in Fig. S11,† the ratio of the NWs in the products increases with the amount of W(CO)6. To further investigate whether tungsten or the carbonyl group impact the final morphology, the products were obtained by replacing W(CO)6 by Mo(CO)6. As shown in Fig. S12,† the shape of the products also resembles ultrathin NWs. This result indicates that the carbonyl group decomposed from the W(CO)6 rather than tungsten acts as the structure-directing agent for the formation of NWs, likely owing to the strong adsorption of carbonyl on Pt atoms. The above results indicate that although HEA NPs could be obtained without the application of W(CO)6, the presence of W(CO)6 is essential for the formation of NWs. The morphology of the product was also affected by CTAB. As shown in the XRD pattern (Fig. S13a†), the diffraction peaks align with Pt (JCPDS No. 04-0802) with a slight positive shift, similar to those of Pt-based HEA NWs. However, the morphology of the products resembles nanoflowers rather than ultrathin NWs (Fig. S13b and c†).
Pt-based electrocatalysts exhibit excellent performance in the HER and MOR.33,34 Herein, the HER activity and stability of HEA NWs were systematically investigated first by a series of electrochemical tests in 1.0 M KOH. The reference electrocatalysts, i.e. Pt/C and HEA NPs, were tested under the same conditions. In this work, all potentials have been referenced to the reversible hydrogen electrode (RHE). As depicted in Fig. 2a, the iR-corrected linear sweep voltammetry (LSV) polarization curves reveal that HEA NWs exhibit a remarkably reduced overpotential compared to the benchmarks (i.e. commercial Pt/C and HEA NPs). At a current density of 10 mA cm−2, HEA NWs exhibit an overpotential of just 14 mV (Fig. 2b, left), a value notably lower than that of HEA NPs (20 mV) and commercial Pt/C (28 mV). Moreover, at an overpotential of 70 mV, the current density for HEA NWs (119.9 mA cm−2) is higher than those of HEA NPs (44.7 mA cm−2) and commercial Pt/C (42.1 mA cm−2) (Fig. 2b, right). As shown in Fig. S14,† among all current densities, HEA NWs demonstrated the lowest overpotential. Since the HEA NWs are designed to reduce the cost of the catalysts, the current densities in the HER also are normalized to the mass of noble metals (Fig. S15†). The overpotential of HEA NWs at 0.5 A mgPGM−1 is only 8 mV, lower than those of HEA NPs (13 mV) and commercial Pt/C (45 mV). At the overpotential of 25 mV, the current density of HEA NWs is 2.0 A mgPGM−1, two and ten times higher than those of HEA NPs and commercial Pt/C, indicating the outstanding mass activity of HEA NWs for the HER. The Tafel slope is used to investigate HER kinetics and mechanisms, and a lower Tafel slope value means more rapid reaction kinetics. As shown in Fig. 2c, the Tafel slope value of HEA NWs is 34.5 mV dec−1, lower than those of HEA NPs (37.6 mV dec−1) and commercial Pt/C (37.9 mV dec−1). The low Tafel slope value suggests that the HER process on HEA NWs follows the Volmer–Tafel mechanism.35 Relative to other reported noble metal-based HER electrocatalysts under alkaline conditions, HEA NWs demonstrate exceptional catalytic activity (Fig. S16†).
Considering the positive correlation between the double-layer capacitance (Cdl) and the electrochemical active surface area (ECSA), cyclic voltammetry (CV) curves of HEA NWs and the reference samples were evaluated across a scan rate range of 20 to 200 mV s−1 (Fig. S17†). Fig. 2d reveals that the calculated Cdl for HEA NWs (29.71 mF cm−2) exceeds those of commercial Pt/C (12.66 mF cm−2) and HEA NPs (17.60 mF cm−2). This result suggests that the HEA NWs, with a large surface area, expose more active sites. Electrochemical impedance spectroscopy (EIS) analysis demonstrates enhanced electron/proton transfer kinetics, with HEA NWs showing the lowest charge transfer resistance (Rct) among all the catalysts (Fig. 2e). The above results highlight the structural advantage of HEA NWs, which expose more active sites and exhibit significantly improved catalytic performance compared to the benchmarks.
Apart from electrocatalytic activity, stability is another crucial factor in assessing the performance of electrocatalysts. To further assess the electrochemical stability of HEA NWs, accelerated degradation tests (ADTs) were conducted, involving 10000 continuous CV cycles. Fig. 2f illustrates the LSV curves of HEA NWs before and after the ADTs. The negligible overpotential shift confirms superior stability versus HEA NPs and Pt/C benchmarks (Fig. S18†). Moreover, continuous operation at 100 mA cm−2 confirms the exceptional durability of HEA NWs, showing negligible degradation over 50 h (Fig. 2f inset). In addition, characterization of the HEA NWs was carried out after the long-term durability test to confirm their structural stability. As shown in Fig. S19,† the change of the shape and the XRD of the NWs is negligible after the i–t test, indicating the structural stability of HEA NWs.
The MOR performance of HEA NWs was also investigated. Before MOR tests, all the catalysts were activated in 1.0 M KOH, which was saturated with N2, through CV measurements. After that, the MOR activities of these catalysts were recorded in 1.0 M KOH + 1.0 M methanol by CV cycles. The HEA NWs exhibit higher activity for the MOR compare with the HEA NPs and commercial Pt/C as displayed in Fig. 3a. Clearly, HEA NWs possess the smallest onset potential (∼531 mV) compared with those of HEA NPs (∼607 mV) and commercial Pt/C (∼741 mV) (Fig. S20†), indicating the MOR can occur more easily on HEA NWs. The currents are normalized to the mass of Pt-group metals (PGM) to compare the mass activities of different catalysts. Specifically, as shown in Fig. 3b, the mass activity of HEA NWs for the MOR is 10.4 A mgPGM−1, which is about 1.9 and 13.0 times as high as that of HEA NPs (5.4 A mgPGM−1) and commercial Pt/C (0.8 A mgPt−1), respectively. Relative to other reported Pt-based electrocatalysts, HEA NWs show remarkable MOR activity in an alkaline environment (Fig. 3c and Table S3†).
![]() | ||
Fig. 3 (a) CV curves of HEA NWs, HEA NPs and commercial Pt/C in 1.0 M KOH + 1.0 M methanol (scan rate: 50 mV s−1). (b) Histograms of mass activities (normalized to the mass of Pt-group metals, PGM) toward the MOR. (c) Comparison of the MOR performance of reported electrocatalysts and HEA NWs.1–12 (d) i–t test of HEA NWs, HEA NPs and commercial Pt/C at 0.72 V vs. RHE. (e) The mass activities on HEA NWs before and after the i–t test. (f) CO-stripping curves of different catalysts. |
Furthermore, the electrocatalytic stability of the catalysts was evaluated using the i–t test. As illustrated in Fig. 3d, HEA NWs maintain superior activity even after 20000 s of i–t testing compared to the other electrocatalysts under the same conditions, suggesting that the HEA NWs possess better electrocatalytic stability. As shown in Fig. S21,† the current densities of HEA NPs and commercial Pt/C decrease 17.9% and 37.1%, respectively, after the i–t test. In comparison, no significant degradation is observed in the current density of HEA NWs (Fig. 3e). And the shape of the NWs after the i–t test is also maintained and the XRD pattern indicates the existence of Pt-based HEA alloys (Fig. S22†). Besides, the XPS of the catalysts was conducted to investigate whether the elements are oxidized after the i–t test for the MOR. As shown in Fig. S23,† the chemical states of all the elements are the same as those before the i–t test. The above characterization demonstrates the structural and compositional stability after the long-term durability test in the MOR. During electrocatalytic reactions, CO intermediates, which arise from the dissociation of alcohol molecules, tend to adsorb onto Pt-based catalysts. This adsorption leads to a decline in the catalyst's activity.34,36,37 To evaluate the CO resistance of the catalysts, as shown in Fig. 3f, the oxidation potential for CO on HEA NWs and HEA NPs is lower than that on commercial Pt/C in the CO stripping curves. This result indicates that the HEA catalysts exhibit excellent CO anti-poisoning performance.
Density functional theory (DFT) calculations are conducted to shed light on the reasons for the improved HER and MOR performance. The theoretical models of HEA (111) and Pt (111) were constructed as shown in Fig. 4a based on the ICP and EDS results. The d-band center, derived from the projected densities of states (PDOSs) in HEA (111) and Pt (111), serves as an effective descriptor for predicting electrocatalytic activity. Relative to pure Pt (111), the d-band center of HEA (111) shifts slightly downward in energy. This shift reduces the over-binding of intermediates, ensuring efficient electrocatalysis (Fig. 4b). In order to thoroughly understand the factors contributing to the enhanced HER activity in alkaline media, detailed calculations of the Gibbs free energy diagrams for the rate-determining step (RDS), i.e. water dissociation, were conducted (Fig. 4c). The corresponding theoretical models are depicted in Fig. S24 and S25.† The HEA possesses a lower activation barrier (ΔGRDS = 0.58 eV) than pure Pt (ΔGRDS = 0.61 eV) for breaking the OH–H bond in H2O, suggesting more favored adsorbed hydrogen (*H) and adsorbed OH group (*OH) production. In addition, as shown in Fig. 4d, the adsorption free energy of *H on HEA (111) (ΔG*H = −0.06 eV) is closer to 0 than that on pure Pt (111) (ΔG*H = −0.40 eV). According to Sabatier's principle, the most active HER electrocatalysts are expected to be positioned at the peak of the volcano plot with ΔG*H = 0, indicating that the HEA exhibits superior intrinsic activity to pure Pt. These results demonstrate that HEA could not only facilitate H2O dissociation but also enhance the *H adsorption during alkaline HER, thus exhibiting improved HER activity.
In the case of the MOR, the free energy diagrams of the full pathway are illustrated in Fig. 4e. The corresponding theoretical models are displayed in Fig. S26 and S27.† Clearly, the energy barriers of all non-spontaneous reaction steps on HEA (111) are lower than those on Pt (111). For example, the energy barrier of the methanol adsorption step on the HEA (111) is 0.15 eV, lower than that (0.24 eV) on Pt (111), indicating the easier adsorption of methanol on Pt-based HEA NWs. As for the initial dehydrogenation steps on HEA (111) and Pt (111), both are spontaneous reaction steps without an energy barrier. Notably, in the whole pathway, the RDS is the oxidation of CHO* to HCOOH. The energy barrier of RDS on HEA (111) is 0.71 eV, lower than that (0.92 eV) on Pt (111), revealing the superior MOR activity of HEA NWs. Fig. 4f illustrates the adsorption energy of CH3OH, CO and CO2 on HEA (111) and Pt (111). The adsorption energy of CH3OH on HEA is smaller than that on Pt, benefiting the following oxidation process of CH3OH. Meanwhile, the adsorption energy of CO2 on HEA (111) is larger than that on Pt (111), indicating that the product (i.e. CO2) of the MOR is easier to desorb from the surface of HEA. Furthermore, the HEA (111) surface demonstrates a reduced tendency to adsorb CO molecules, thereby enhancing its tolerance to CO-induced poisoning in the MOR. The optimized electronic structure of HEA promotes stronger binding of reaction intermediates while facilitating the RDS by reducing its activation energy in alkaline HER and MOR processes.
Footnotes |
† Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d5ta03751d |
‡ These authors contributed equally to this work. |
This journal is © The Royal Society of Chemistry 2025 |