Long
Chen‡
a,
Xiying
Jian‡
a,
Qingsheng
Gao
b,
Zhijie
Cao
c,
Zibo
Chen
*d and
Huai-Jun
Lin
*a
aInstitute of Advanced Wear & Corrosion Resistance and Functional Materials, Jinan University, Guangzhou 510632, PR China. E-mail: hjlin@jnu.edu.cn
bDepartment of Chemistry, Jinan University, Guangzhou, 510632, PR China
cCollege of Physics, Ningxia University, Yinchuan, Ningxia 750021, PR China
dInstitute of Chemical & Process Engineering, University of Strathclyde, Glasgow G4 0JQ, UK. E-mail: zibo.chen@strath.ac.uk
First published on 26th March 2025
Designing bimetallic alloys and providing dual active sites is an effective way to achieve efficient alkaline hydrogen evolution reaction (HER). However, the preparation and determination of an optimal composition remain significant challenges for immiscible alloys. In this study, the amorphous Cu–Mo alloys with a wide composition range were successfully prepared by magnetron sputtering. Among these alloys, the amorphous Cu50Mo50 alloy demonstrates excellent alkaline HER activity with low overpotentials of 57 and 149 mV at 10 and 100 mA cm−2, respectively. Moreover, it demonstrates outstanding long-term stability at 300 mA cm−2. The results demonstrate that an electronic interaction exists between the Cu and Mo atoms in the amorphous Cu50Mo50 alloy, where the Cu and Mo act as adsorption sites for OH and H intermediates, respectively. Furthermore, the amorphous Cu50Mo50 alloy also exhibits favorable intermediates adsorption and water dissociation abilities, which facilitate the alkaline HER process. The research provides a novel insight into the rational design and preparation of advanced alkaline HER catalysts.
Mo-based materials have garnered significant interest as active species for HER.12–15 However, exploring efficient Mo-based catalysts in HER while achieving long-term stability still faces enormous challenges.16 Alloying is an efficient method of improving catalyst performance due to the ligand effect caused by the formation of heteroatoms and the strain effect caused by the change in bond length.17 In a preceding study, the Brewer–Engel valence bond theory was proposed as a guiding principle for the design of advanced HER catalysts.18 According to the theory, combining the early transition metals with empty or partially filled d-orbitals and the late transition metals with internal paired d-orbitals will result in the remarkable synergistic effect of electrocatalysis.19 For instance, in the report by Miljanić et al., alloys including Pt2Mo and Hf2Fe have been employed as efficient alkaline HER catalysts.20
Cu, being a low-cost late transition metal, has attracted considerable interest in recent years, particularly within the field of catalysis.21 However, the weak hydrogen adsorption properties of Cu have resulted in a paucity of reports on Cu-based catalysts used for HER.22 In the study by Li and colleagues, the significant role of Cu in the cleaving of H–O bonds in water molecules and the absorption of OH is impressive.23 In non-acidic conditions, the rate-determining step in HER is usually the sluggish water dissociation process (H2O → H* + OH*). The high affinity of catalysts for the hydroxyl intermediate (OH*) can facilitate enhanced reaction rates.24 In light of the aforementioned considerations, alloying Mo with Cu could result in unprecedented alkaline HER performance.
Amorphous materials exhibit long-range disorder structures, which are very different from the crystalline materials.25 The structural flexibility, adjustable composition and abundance of active sites in amorphous materials render them highly promising as catalysts for HER.26 Specifically, amorphous catalysts display the following advantages in comparison to their crystalline counterparts: (i) in comparison to crystalline materials with a fixed coordination environment, amorphous materials can be accurately modified at the atomic ratio, which greatly facilitates the design of optimal catalysts with an appropriate elemental ratio; (ii) the increased diffusion of electrolyte into the internal regions of amorphous matrices facilitates the formation of a larger electrolyte–electrocatalyst interface; (iii) amorphous structure is rich in dangling bonds or coordination unsaturated atoms, which may provide more active sites to promote electrochemical processes.27,28
The combination of Cu and Mo has been demonstrated in several reports to enhance HER activity.29,30 However, these reports exclusively consider Cu–Mo alloys with a highly constrained composition range. The activity of an electrocatalyst is significantly affected by its composition.31 The preparation of amorphous Cu–Mo alloys allow for the effective improvement of the catalytic activity, through the alteration of the composition. It is worthy of note that the Cu–Mo system is immiscible in equilibrium state (Fig. S1†), thus the preparation of Cu–Mo alloys by conventional metallurgy methods remains challenge. The magnetron sputtering technique allows for the direct deposition of a multi-metal catalyst onto a substrate, while simultaneously enabling the alteration of the catalyst's composition and loading.32 Thus, it is feasible to prepare amorphous Cu–Mo alloys for efficient HER via magnetron sputtering.
In this work, we creatively obtained a series of Cu–Mo alloys with a wide composition range in the immiscible Cu–Mo system by magnetron sputtering, and employed them as HER catalysts. Among them, the prepared amorphous Cu50Mo50 exhibits optimal HER performance with a low overpotential of 57 mV at 10 mA cm−2 and robust long-term stability at 300 mA cm−2 in 1.0 M KOH. The final results in this paper prove the feasibility of preparing highly effective alkaline HER catalysts by simple synthesis method and balancing dual active components. This synthesis method and design idea would create more opportunities for developing advanced alkaline HER catalysts.
The phase information of the as-prepared CuxMoy alloys was examined by X-ray diffraction (XRD) analysis, as shown in Fig. 1a. The CuxMoy alloy exhibits an amorphous structure within the atomic ratio range of approximately 43:
57 to 75
:
25, as evidenced by the presence of weak and broad X-ray diffraction peaks at around 40°. The amorphization of the Cu–Mo alloys rise and then fall, as the Cu content increases. The strong XRD peaks of the Cu film are attributed to Cu (111) and Cu (200) planes (PDF# 04-0836). The strong diffraction peaks observed in the Mo film are attributed to Mo (110) planes (PDF# 42-1120). It can be observed that the diffraction peaks of the crystalline CuxMoy apparently shift to higher angles with increasing Cu content, suggesting lattice contraction caused by the smaller atomic radius of the Cu atom (128 pm) compared to that of Mo (139 pm).30
The surface morphology of the Cu50Mo50 alloy is shown in Fig. 1b, as obtained by scanning electron microscopy (SEM). The surface of the Cu50Mo50 alloy exhibits small island microstructures, a common structure observed during the preparation of alloys by magnetron sputtering.33 Energy Dispersive X-ray spectroscopy (EDS) elemental mapping images exhibit that Cu and Mo are uniformly distributed on the surface of the Cu50Mo50 alloy (Fig. S2†). The results of the cross-sectional EDS images demonstrate that Ni does not diffuse into the Cu50Mo50 alloy (Fig. S3†). Fig. 1c presents a transmission electron microscopy (TEM) image of the Cu50Mo50 alloy that has been thinned by focused ion beam technology (FIB). We then carried out selected area electron diffraction analysis on the film sample (Fig. 1c and d). The halo ring observed in the selected area electron diffraction pattern (SAED, Fig. 1d) provides further evidence of the formation of amorphous structures in the Cu50Mo50 alloy. Fig. 1e together with the inset are high-resolution transmission electron microscopy (HRTEM) image and the corresponding fast Fourier transforms (FFT) image of the Cu50Mo50 alloy. The absence of discernible periodic lattice stripes in the HRTEM image and the appearance of diffusion ring in the FFT image illustrate the amorphous structure of the Cu50Mo50 alloy.
X-ray photoelectron spectroscopy (XPS) was employed to characterize the valence state information. The XPS survey spectra demonstrate the presence of Cu and Mo elements in the CuxMoy alloys (Fig. 2a). Fig. 2b shows the high-resolution Cu 2p spectra. For the Cu sample, the peaks located at 932.38 and 952.18 eV are correspond to the Cu0 2p3/2 and Cu0 2p1/2, while the signals located at 934.73 and 954.33 eV correspond to the Cu 2p3/2 and Cu 2p1/2 of Cu(OH)2 (Cu2+). The signals located at 943.83, 946.65 and 962.37 eV are assigned to the satellite peaks of Cu2+. For the CuxMoy alloys, the peaks located at 932.2–932.6 eV are assigned to Cu 2p3/2, and the peaks located at 952.1–952.4 eV are assigned to Cu 2p1/2, which reveal the existence of Cu+ peaks. The other peaks observed at 943.5–946.8 eV are assigned to the satellite peaks of Cu+. It is worth noting that all the characteristic peaks of Cu 2p in Cu85Mo15 and Cu50Mo50 alloys are shifted to lower binding energies compared to Cu28Mo72 alloy, indicating an increased electron density of Cu atoms. Fig. 2c displays the Cu LMM spectra to assist in the analysis, as the XPS peak positions of Cu+ and Cu are indistinguishable. The Cu LMM peak in the range of 918–919 eV is classified as the signal of Cu0. However, it is difficult to analyze the valence state using the Cu LMM due to the proximity of the peaks of Cu2O and Cu(OH)2 in the range of 916–917 eV. The presence of Cu(OH)2 peaks in the XPS spectra of Cu 2p (Fig. 2b) indicates that the LMM peak in the region of 916–917 eV is Cu(OH)2 for the Cu sample and Cu2O for the CuxMoy samples (Fig. 2c).
![]() | ||
Fig. 2 XPS characterization of the CuxMoy alloys. (a) Survey spectra. (b) Cu 2p. (c) Cu LMM. (d) Mo 3d. (e) Valance band spectra. (f) O 1s. |
Fig. 2d illustrates that the XPS spectra of Mo 3d shows five spin–orbit doublets. For the CuxMoy alloys, the first doublet at 227.9–228.3 and 231.1–231.5 eV is attributed to Mo0, while the second doublet at 228.7–228.9 and 231.9–232.1 eV is attributed to Mo2+. The third doublet at 229.7–230.0 and 232.8–232.2 eV corresponds to Mo4+, while the fourth doublet at 231.6–231.8 and 234.8–235.0 eV corresponds to Mo5+. Moreover, the fifth doublet at 233.6–233.7 and 236.7–236.9 eV is assigned to Mo6+. For the Mo sample, the peaks located at 228.0 and 230. 95 eV are assigned to Mo 3d5/2 and Mo 3d3/2 of Mo0, while the peaks located at 228.51 and 231.61 eV correspond to Mo 3d5/2 and Mo 3d3/2 of Mo2+. The peaks located at 229.48, 232.53, 231.43 and 234.58 eV are assigned to Mo4+ 3d5/2, Mo4+ 3d3/2, Mo5+ 3d5/2 and Mo5+ 3d3/2, respectively. Additionally, the peaks located at 233.28 and 236.38 eV are attributed to Mo6+. After alloying, the characteristic peaks of Mo 3d in the CuxMoy alloys all shift to higher binding energies than the Mo sample, indicating a decrease in electron density of the Mo atoms (Fig. 2d). This XPS peak shift phenomenon can be attributed to the electron transfer between Cu and Mo. The electronic interaction between Cu and Mo can be further demonstrated by the difference in electronegativity between Cu (1.9) and Mo (1.8).30 The electronic structure of Cu, Cu85Mo15, Cu50Mo50, Cu28Mo72 and Mo samples is further analyzed by valence band spectra (Fig. 2e). The high density of states (DOS) near the Fermi level exhibited by the Cu85Mo15 and Cu50Mo50 alloys demonstrates higher carrier concentration, which indicates a modification of the electronic structure. Additionally, the XPS spectra of O 1s were also examined. There are three visible peaks in the O 1s XPS spectra of all samples. For the Cu50Mo50 alloy, the three peaks observed at 530.28, 531.43 and 532.80 eV are attributed to metal–oxygen bonds (lattice oxygen), hydroxyl and adsorbed H2O, respectively (Fig. 2f).
Additionally, the HER performances of the CuxMoy alloys in acidic media were also investigated. Fig. S9a† shows that the CuxMoy alloys with an atomic ratio of 28:
72 to 56
:
44 exhibit similar LSV curves necessitating an overpotential of approximately 120 and 214 mV to drive 10 and 100 mA cm−2, respectively. Regarding the catalytic activity of acidic HER, it can be observed that the Mo sample exhibits higher activity than Cu, while Ni foam shows the lowest activity (Fig. S10†). As illustrated in Fig. S9b,† the Tafel slope of the Cu50Mo50 alloy in an acidic environment is 104.1 mV dec−1, a value that is larger than its 85.6 mV dec−1 in an alkaline environment (Fig. 3d). This indicates that the HER process of the Cu50Mo50 alloy is slower in the acidic media than in alkaline media. All CuxMoy alloys follow the Volmer–Heyrovsky step in acidic media. As the Cu content in the CuxMoy alloys increases, the Tafel slope decreases, indicating that the kinetics of hydrogen evolution is gradually improved (Fig. S9b†). The charge transfer capability of the CuxMoy alloys was analyzed by EIS. The Nyquist plots of the CuxMoy alloys in an 0.5 M H2SO4 exhibit two semicircles (Fig. S9c†). The Rct values of the Cu43Mo57, Cu50Mo50 and Cu56Mo44 alloys are 52.67, 49.48 and 51.61 Ω, respectively, which are smaller than those of the Cu28Mo72 (62.24 Ω) and Cu85Mo15 (62.11 Ω) alloys, implying faster charge transfer (Fig. S9c and Table S3†). The detailed fitting process of the Nyquist plot of the CuxMoy alloys in 0.5 M H2SO4 is shown in Fig. S11.† As illustrated in Fig. S9d and S12,† the Cdl values of Cu28Mo72, Cu43Mo57, Cu50Mo50 and Cu56Mo44 are 64.2, 69.2, 68.1 and 62.1 mF cm−2, respectively. These values indicate that the CuxMoy alloys with an atomic ratio of 28
:
72 to 56
:
44 exhibit more exposed active sites in comparison to the Cu85Mo15 alloy (18.5 mF cm−2).
HER catalysts for industrial applications require long-term stability at high current densities. Consequently, large current density stability tests were conducted on the Cu50Mo50 alloy. As illustrated in Fig. 4b, the Cu50Mo50 alloy exhibits outstanding stability during the 300 mA cm−2i–t test, with a current density retention rate of 88.1%. Furthermore, there is an almost overlapping LSV curve before and after the stability test (Fig. S14†). Following the 12 hours i–t test at 500 mA cm−2, the current retention rate of the Cu50Mo50 alloy is 72.9%. In the following section, we will discuss the reasons for the deterioration of stability in the 500 mA cm−2i–t test. To explore the reasons for the enhancement in the stability of the Cu50Mo50 alloy in the 300 mA cm−2i–t test, an analysis of the Bode phase plots was conducted under a range of applied potentials. Fig. 4c illustrates that as the applied potential is increased, the amorphous Cu50Mo50 alloy displays an accelerated Heyrovsky step, as evidenced by the reduction in phase angle within the mid-frequency range in the Bode phase plots. The middle-frequency range (MF: 100–103 Hz) in Bode phase plots is thought to correspond to the electrochemical response of charge transfer.38 Moreover, the phase angle within the mid-frequency range of Bode phase plots reflects the rate of Heyrovsky step.39
In light of the preceding analysis, we proceeded to investigate the long-term stability of the Cu50Mo50 alloy. As illustrated in Fig. 4d, a long-term i–t test at 300 mA cm−2 is conducted on Cu50Mo50 in 1.0 M KOH. The current density does not exhibit a continuous attenuation during the i–t test, and the Cu50Mo50 alloy demonstrates a reduction in overpotential following the test (The inset in Fig. 4d). In Section 2.4, we will discuss the phenomenon of overpotential reduction after long-term i–t test. Furthermore, the durability of the sample is evaluated using a chronopotentiometry (e–t) test (Fig. 4e). The results indicate that the overpotential increases by a mere 67 mV after applying a constant current density of 300 for 150 hours. The above findings demonstrate that the amorphous Cu50Mo50 alloy exhibits remarkable high current density and long-term stability in alkaline HER process.
![]() | ||
Fig. 5 XPS spectra of the Cu50Mo50 alloy after the 12 h i–t test at 100 mA cm−2. (a) Survey spectra. (b) Cu 2p. (c) Cu LMM. (d) Mo 3d. (e) O 1s. |
The amorphous Cu50Mo50 alloy exhibits remarkable stability at 300 mA cm−2 in 1.0 M KOH (Fig. 4b). SEM and TEM were employed to gain insights into the morphology and composition of the amorphous Cu50Mo50 alloy after the i–t test at 300 mA cm−2. As illustrated in Fig. 6a, the surface of the Cu50Mo50 alloy exhibits sharp particles, which differ from the morphology observed in the original sample. After the stability test, the distribution of the Cu and Mo elements on the surface of the Cu50Mo50 alloy is observed to remain uniform (Fig. S15a–c†). The cross-section EDS mapping indicates that the presence of sharp particles on the surface of the sample after the stability test can be attributed to the corrosion product Cu(OH)2 (Fig. S16†). As shown in Fig. 4d, after long-term i–t test, the sample showed an increasing chronoamperometric curve and a decreasing overpotential. This prompted us to conduct relevant characterization and tests to explore this phenomenon. The SEM images of the surface of the Cu50Mo50 alloy were obtained following i–t tests at 300 mA cm−2 over different periods (Fig. S17†). It can be found that after long-term i–t test, there are many holes on the surface of Cu50Mo50 alloy (Fig. S17b†), which is different from that of 12 h. The existence of holes increases the contact between catalyst and electrolyte, thus increasing active sites and improving catalytic performance. We subsequently obtained the Cdl prior to and following the long-term i–t test to assess the electrochemical active area (Fig. S18†). The Cdl is obtained by testing the CV at different scanning speeds in the non-Faraday region (Fig. S18b and c†). The Cdl values indicate that the electrochemical active area of the sample increases after long-term i–t test, which is consistent with the results reflected by SEM images.
Moreover, the morphologies of the Cu50Mo50 alloy were obtained after 12 hours i–t tests at 100 and 500 mA cm−2 (Fig. S19†). It can be found that under the condition of 500 mA cm−2, the surface of the sample is covered with large-scale corrosion products (Fig. S19b†). The presence of this surface corrosion product hinders the contact between the catalyst and electrolyte, which deteriorates the catalytic performance. We carried out Inductively Coupled Plasma technology (ICP) tests to analyze the element content in the electrolyte after the i–t test (Table S4†). As shown in the ICP results, the electrolyte contains the highest amount of Mo at 300 mA cm−2, and the highest amount of Cu at 500 mA cm−2. When considering the stability results (Fig. 4b), it can be concluded that the factors affecting the stability of the Cu50Mo50 alloy are not only film shedding or element dissolution. It is also important to consider the impact of corrosion products on the performance of catalysts (Fig. S19†).
Fig. 6b presents a low-magnification TEM image of the Cu50Mo50 alloy after the stability test. As illustrated in Fig. 6c, the inconspicuous diffraction spots in the SAED pattern indicate that the evolution from amorphous to crystalline in the Cu50Mo50 alloy occurs during the 12 h i–t test at 300 mA cm−2. The presence of the diffuse ring in SAED pattern shows that the amorphous structure remains in the Cu50Mo50 alloy. The amorphous structure is thermodynamically metastable, and the samples prepared by magnetron sputtering are in a non-equilibrium state. The intense hydrogen evolution reaction gradually leads to the appearance of crystalline phases in the amorphous structure. The diffraction rings are assigned to the (103) plane in MoO3, the (113) plane in Cu2O and the (022) plane in MoO2 (Fig. 6c). The results in SEAD pattern demonstrate that amorphous and crystalline phases coexist in the Cu50Mo50 alloy after the stability test. The blurred lattice fringes observed in the HRTEM image and the diffused halo in the FFT pattern further illustrate the retained amorphous structure of the Cu50Mo50 alloy after the stability test (Fig. 6d and e).
It is important to note that the influence of OH on the catalytic reaction rate cannot be ignored in alkaline HER. Chen et al. verified through experiments and simulations that the adsorbed OH (OHad) plays a significant role in alkaline HER, regulating water connectivity and improving the hydrogen bonding network in the electric double layer (EDL) region.41 Therefore, the OH adsorption capacity of the CuxMoy alloys was analyzed (Fig. 7b). The Cu50Mo50 alloy demonstrate a more pronounced reversibility in the adsorption and desorption of OH compared to other CuxMoy alloys. Compared with the Cu85Mo15 alloy, the smaller range of current density of the Cu50Mo50 alloy can be attributed to the rapid OH transfer process.42 In comparison to Mo, Cu exhibits a strong OH adsorption peak, and with the Cu content increased, the OH adsorption of the CuxMoy alloys displays a gradual increase (Fig. 7b).
Raman spectra were obtained for Cu and Mo at the potential of 0.85 V (V vs. RHE). The Raman spectrum of Cu exhibits a peak at 495 cm−1, indicating a strong adsorption of OH by Cu (Fig. S20†).43 No obvious peak was observed in the Raman spectra of Mo (Fig. S20†). Accordingly, in the alkaline HER process, Cu atoms in the CuxMoy alloys work as the active sites for OH adsorption. Meanwhile, Mo atoms exhibiting high HER activity function as sites for the conversion of hydrogen. The synergistic effect of adjacent atoms significantly enhances the alkaline HER process, as isolated Cu and Mo atoms are unable to effectively promote the overall electrocatalysis process. Notably, excessive adsorption of OH will result in the poisoning of the active site of the catalysts, while insufficient adsorption will reduce the water dissociation ability, thus affecting the HER efficiency.44 Therefore, the appropriate OH adsorption/desorption capacity of the catalyst is conducive to the alkaline HER process. This may be the main reason for the excellent overall alkaline HER performance of the amorphous Cu50Mo50 alloy. Due to the favorable intermediate adsorption, there is no discernible signal in the in situ Raman spectrum of the amorphous Cu50Mo50 alloy (Fig. S21†).
Furthermore, the interface water structure of the Cu50Mo50 alloy was investigated through in situ Raman spectroscopy in 1.0 M KOH (Fig. 7c). We decomposed the interfacial water peak into three peaks. The water peaks at approximately 3230 and 3440 cm−1 are assigned to tetrahedral (ice-like water) and trihedral ligand water (liquid-like water), respectively.13 The peak at approximately 3610 cm−1 is indicative of the presence of dangling OH bonds of interfacial water, representing the free water molecules that are readily transported. The activation energies of water dissociation increased in the following sequence: the free water, the liquid-like water, and the ice-like water.45 To gain a more intuitive understanding of the distribution of interface water, a Gaussian fitting is applied to the Raman signal of the interface water (Fig. S22†). As illustrated in Fig. S22 and S23a,† the strength of free water decreases gradually with an increase in applied potential, which proves the effective dissociation of water on the Cu50Mo50 alloy.46 In comparison with unitary Cu and Mo, the free water strength of the Cu50Mo50 alloy demonstrates a more significant weakening within the applied potential range, thereby indicating its optimized water dissociation ability (Fig. S23†).
The synergistic effect of Mo and Cu is not only evident in the adsorption of reactants, but also in the electronic interaction between them. As an electron acceptor, Cu in the Cu50Mo50 alloy exhibits an elevated electron density, which reduces the energy barrier for the adsorption of OH (Fig. 2b). This is due to the enhanced ease with which electrons can be transferred to the O 2p orbital that adsorb OH.47 As an electron donor, Mo in the Cu50Mo50 alloy shows more positive charge, which is thought to result in a weakened Mo–H bond (Fig. 2d).48 The Mo–H bond energy is strong, and the reduction of electron filling to Mo–H can result in a weakening of the bond.49 Furthermore, the enriched DOS near the Fermi level exhibited by the Cu50Mo50 alloy is thought to be conductive to the HER process (Fig. 2e).50 The aforementioned discussion reveals the alkaline HER mechanism of the amorphous Cu50Mo50 alloy (Fig. 7d). The combination of Mo and Cu results in the effective dissociation of water molecules, with Cu acting as an adsorption site for OH and Mo acting as an adsorption site for H. Subsequently, the H* species are effectively desorbed at the Mo sites, thus producing H2 as a result. Concurrently, the OH* species on the Cu sites transferred effectively for the subsequent reaction. This explains that the alkaline HER performance of the CuxMoy alloys is more dependent on the composition than the acidic HER performance (Fig. 3a and S9a†). Furthermore, the long-range disordered structure, small grain size and high surface area of amorphous materials enhance the dual-site synergistic effect and atom utilization rate in the amorphous Cu50Mo50 alloy. Therefore, by preparing amorphous bimetallic alloys and regulating functional components, it is expected to achieve efficient water dissociation and balanced intermediate adsorption, thus promoting the alkaline HER process.
(1) The amorphous CuxMoy alloys with a wide composition are prepared by magnetron sputtering for HER, a method which can be extended to the preparation of other immiscible systems catalysts for HER applications.
(2) The amorphous Cu50Mo50 catalyst exhibits a considerably low overpotential of 57 mV to reach 10 mA cm−2 and outstanding long-term stability at a large current density of 300 mA cm−2 in 1.0 M KOH. Furthermore, excellent activity with an overpotential of 120 mV at η10 and robust stability is observed in 0.5 M H2SO4 as well.
(3) Benefiting from the synergistic effect of dual active sites, the amorphous Cu50Mo50 alloy with modified electronic structure shows optimized water dissociation and good intermediate adsorption capacity, thus accelerating the alkaline HER process.
This work offers a viable approach for the preparation of immiscible alloys with high HER activity, while also providing valuable insights for the design of efficient alkaline HER catalysts.
Footnotes |
† Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d5ta00701a |
‡ Both authors contributed equally to this work. |
This journal is © The Royal Society of Chemistry 2025 |