Yiqing
Sun
ab,
Yiwei
Bao
ab,
Di
Yin
c,
Xiuming
Bu
*b,
Yuxuan
Zhang
c,
Kaihang
Yue
b,
Xiaoshuang
Qi
b,
Ziyan
Cai
b,
Yongqiang
Li
d,
Xiulan
Hu
*a,
Johnny C.
Ho
*ce and
Xianying
Wang
*b
aCollege of Materials Science and Engineering, Nanjing Tech University, Puzhu South Road No. 30, 211816, Nanjing, Jiangsu, China. E-mail: whoxiulan@163.com
bCAS Key Laboratory of Materials for Energy Conversion, Shanghai Institute of Ceramics, Chinese Academy of Sciences, Shanghai 200050, P. R. China. E-mail: buxiuming@mail.sic.ac.cn; wangxianying@mail.sic.ac.cn
cDepartment of Materials Science and Engineering, City University of Hong Kong, Kowloon, Hong Kong SAR 999077, P. R. China. E-mail: johnnyho@cityu.edu.hk
dJoint Laboratory of Graphene Materials and Applications, Shanghai Institute of Microsystem and Information Technology, Chinese Academy of Science, Shanghai 200050, P. R. China
eInstitute for Materials Chemistry and Engineering, Kyushu University, Fukuoka 816-8580, Japan
First published on 29th March 2024
Searching for a stable and efficient electrocatalyst for the hydrogen evolution reaction is still challenging, especially under a wider pH operation condition. In this study, a multicomponent Ni17W3/MoO3−x/WO3−x catalyst was designed and synthesized, in which the unique hierarchical structure of entangled nanorods confined in a polyhedral framework ensures the maximum utilization of active sites. Significantly, electrochemical performance can be regulated by adjusting the oxygen vacancy concentration of the metal support. Combined with various characterization techniques, we discovered that abundant oxygen vacancies in the MoO3−x/WO3−x support not only significantly enhanced the hydrogen insertion/extraction kinetics in the metal oxide but also increased the hydration capacity, resulting in an efficient hydrogen adsorption/transfer/desorption kinetics on the Ni17W3/MoO3−x/WO3−x surface and interface. As a result, the fabricated electrocatalyst exhibits an ultralow overpotential of 16, 42, and 14 mV at 10 mA cm−2 in alkaline, neutral, and acid electrolytes, respectively. Our work proves the important role of metal oxide supports in the hydrogen spillover process.
The construction of metal/metal oxides has been proven to be an efficient strategy since a metal oxide support modulates an active metal phase's hydrogen intermediate (H*) adsorption/desorption behavior through strong metal–support interactions (SMSIs).10–12 Among various metal oxide support candidates, WOx has attracted widespread attention because of its excellent electronic conductivity and outstanding synergistic effect with active metals.13 Meanwhile, the hydrogen spillover phenomenon is observed on the metal/WOx interface in two steps. First, H* transfers from the metal (ΔGH-metal < 0) to the WOx surface (ΔGH-support > 0),14 causing the active metal surface to rapidly re-expose. Then, H* inserts/extracts into the WOx lattice to provide additional active sites for the hydrogenation of WOx. Previous studies have primarily focused on reducing the hydrogen spillover transfer barrier through active metal modification (EG–Pt/CoP and Pt SA/WOx).15,16 To the best of our knowledge, few studies have investigated the effect of support modification strategy on the HER performance.
In this work, we designed and synthesized Ni17W3/MoO3−x/WO3−x multicomponent catalysts via an in situ confinement method, in which Ni17W3 functions as the active metal phase and MoO3−x/WO3−x serves as the support. Moreover, the fabricated catalyst's oxygen vacancy concentration can be adjusted simply by changing the mole ratio between MoO3 and WO3. Next, by tracking phase evolution and mass/charge transfer kinetics with in situ Raman and in situ electrochemical impedance spectroscopy techniques, the superior performance of the catalyst results from the abundant oxygen vacancies of the metal oxide support, which accelerates phase transition from the metal oxide to hydrogen bronze, further resulting in a rapid H* adsorption/transfer/desorption kinetics on the Ni17W3/MoO3−x/WO3−x surface and interface. As a result, the fabricated Ni17W3/MoO3−x/WO3−x catalyst shows superior HER performance at all pH values, requiring only 14, 42, and 16 mV overpotential to reach a geometric current density of 10 mA cm−2 in acid, neutral, and alkaline electrolytes, respectively. Our research confirms the practical potential of Ni17W3/MoO3−x/WO3−x and provides new design concepts for metal/metal oxide catalysts.
Further detailed structural information is explored by high-resolution TEM (HRTEM). The characteristic lattice fringes of 0.211, 0.260, and 0.386 nm correspond well to the (111) plane of Ni17W3, (220) plane of WO3 and (011) plane of MoO3, respectively (Fig. 1g). In addition, the characteristic lattice fringes of 0.210, 0.260, and 0.340 nm in NWP catalysts correspond to the (111) plane of Ni17W3, (220) plane of WO3, and (−110) plane of WO2, respectively (Fig. S6†).
Combined with the crystal and phase information obtained through the X-ray diffraction (XRD) patterns (Fig. S7–S9†), the characteristic peaks of NWP located at 23.48°, 33.60°, 41.40° and 48.04° correspond to the (020), (202), (222) and (040) facets of WO3 (JCPDS 46-1096), and the peaks at 25.79° and 37.07° are indexed to the (011) and (−211) planes of WO2 (JCPDS 05-0431).18 Also, three peaks at 43.70°, 50.88°, and 74.82° correspond to the diffraction of (111), (200) and (220) planes of the Ni17W3 phase, respectively (JCPDS 65-4828).19 Thus, the crystal phase composition for NWP is Ni17W3/WO2/WO3. On the other hand, for the NW1M0.5R, NW1M1R, and NW1M1.5R samples, the characteristic peaks for WO2 disappear totally, and new peaks located at 23.05°, 25.30°, 36.09°, and 47.28° appear, which can be assigned to MoO3 (JCPDS 89-1554).20 Thus, the crystal phase composition for these three samples is Ni17W3/MoO3/WO3. The characteristic NMR peak at 50.9° corresponds to the (310) plane of Ni4Mo (JCPDS 03-1036). The energy dispersive X-ray (EDX) elemental mapping of NW1M1R (Fig. 1h) shows that Mo, Ni, and W are present and evenly distributed throughout the entire nanorod, indicating the successful preparation of Ni17W3/MoO3/WO3 ternary heterostructures.
The surface chemical composition and electronic structure of the obtained electrocatalysts are then assessed by X-ray photoelectron spectroscopy (XPS). Obviously, the obtained spectrum confirms the existence of Ni, Mo, W, and O in NW1M0.5R, NW1M1R, NW1M1.5R (Fig. S10†). The high-resolution Ni 2p spectrum in Fig. S11† shows two peaks at ∼852.8 and ∼869.2 eV attributed to metallic state Ni0, while the other four peaks are corresponding to Ni2+ (∼856.1 and ∼873.7 eV) and Ni3+ (∼857.6 and ∼875 eV) and their corresponding satellite peaks (∼861.7 and ∼880.1 eV).21 The high-resolution W 4f spectrum (Fig. S12†) of NW1M1R displays three doublets, which can be attributed to W0 (∼31.4 and ∼33.6 eV), W4+ (∼32.8 and ∼33.9 eV) and W6+ (∼35.5 and ∼37.5 eV), respectively.22,23 In addition, the Mo 3d spectrum (Fig. S13†) exhibits three sets of peaks, corresponding to Mo4+ (∼229.4 and ∼231.7 eV), Mo5+ (∼230.6 and ∼234.3 eV) and Mo6+ (∼232.7 and ∼235.5 eV), respectively.24,25 It is also noted that after Mo introduction, the Ni 2p and W 4f binding energies of NW1M1R are the largest compared with the NWP, blue-shifted by 0.3 and 0.6 eV. For the Mo 3d spectrum, no obvious binding energy shift is observed.26,27 Moreover, the O 1s XPS spectrum shows three typical peaks centered at 530.4, 531.5, and 532.6 eV, corresponding to the lattice oxygen, oxygen vacancy, and absorbed oxygen, and NW1M1R shows the highest oxygen vacancy concentrations (Fig. S14†).28,29 To further confirm the presence of oxygen vacancies, the prepared catalysts were subjected to electron paramagnetic resonance (EPR) spectroscopy. NW1M0.5R, NW1M1R and NW1M1.5R all show EPR signals at g = 2.002, while NWP shows a signal at g = 2.003.30,31 The EPR results further support that NW1M1R exhibits a stronger EPR signal than NW1M0.5R, NW1M1.5R and NWP (Fig. S15†). When the W and Mo atomic ratio is 1:
1, the sample exhibits the highest oxygen vacancy concentration, which can be attributed to the lattice mismatch between cubic phase α-WO3 and monoclinic phase β-MoO3. The above electronic structure analysis results indicate that the optimized metal oxide ratio can not only adjust the oxygen vacancy concentration but also enhance the interface contact among these material phases.32
To shed light on the excellent electrochemical performance of the fabricated electrocatalyst, we first investigated the HER activity of each catalyst in 1.0 M KOH via linear scanning voltammetry (LSV). As shown in Fig. 2a and S14,† among all the tested samples, NW1M1R establishes the best HER catalytic performance, needing a lowest overpotential of only 16 mV to achieve the current density of 10 mA cm−2, which is significantly superior than those of NWP (72 mV), NW1M1.5R (44 mV), NW1M0.5R (54 mV), NMR (51 mV) and Pt/C (29 mV), respectively. Then, the corresponding Tafel slope of NW1M1R is fitted to 34.9 mV dec−1, superior to those of NWP (72.4 mV dec−1), NW1M1.5R (54.8 mV dec−1), NW1M0.5R (62 mV dec−1), NMR (54.1 mV dec−1) and commercial Pt/C (49.2 mV dec−1) (Fig. 2b and S16†), indicating a fast HER kinetics.33 Specifically, the Tafel slope value of NW1M1R is close to 30 mV dec−1, suggesting that the HER process follows the Volmer–Tafel mechanism, where hydrogen is formed by the combination of neighboring adsorbed H* rather than by the combination of H* and water molecules in the electrolyte. In addition, electrochemical impedance spectroscopy (EIS) was further employed to analyze the HER kinetics at the interface between the electrode and electrolyte (Fig. 2d). The Nyquist plots show that NW1M1R has the smallest radius, while the equivalent circuit diagram (Fig. S17†) yields that NW1M1R also has the lowest charge transfer resistance (Rct = 2.8 Ω). Notably, compared to the recently reported HER catalysts, the excellent HER performance of NW1M1R is almost the best among them (Fig. 2c and Table S1†).
After that, double-layered capacitances (Cdl) are measured by cyclic voltammetry (CV) at different sweep rates from 20 to 100 mV s−1 to assess the electrochemical active surface area (ECSA) of each catalyst (Fig. S18 and S19†). It is observed that the Cdl value of NW1M1R is 22.24 mF cm−2, which is 4.82 times higher than that of NWP (4.61 mF cm−2). When the current densities are normalized with respect to ECSA, NW1M1R still exhibits a much larger current density than other samples during the entire voltage range, indicating that NW1M1R is indeed intrinsically more active than NWP, NW1M1.5R, and NW1M0.5R (Fig. S20†). In addition to the high catalytic performance demonstrated, the Faraday efficiency of the NW1M1R was determined to be 99.37%, nearly 100% (Fig. S21†). The long-term HER performance of NW1M1R at different current densities and 40 hours for each current density (160 hours in total) and at a high current density of 500 mA cm−2 (100 hours) shows that the voltage has no apparent upward trend (Fig. 2e and S22†), indicating that this unique hierarchical structure improves the stability by preventing aggregation, which is usually one of the main factors leading to catalyst deactivation. The almost unchanged NW1M1R nanostructure, even after 160 h of stability testing, further supports the superiority of the hierarchical structure during the HER process (Fig. S23 and S24†). After long-term stability, the sample was also subjected to XPS characterization (Fig. S25–S27†). The comparison of the pre- and post-reaction states showed that Ni, Mo, and W promote the HER process in the form of high valence states. In addition, because of the superior HER performance under alkaline conditions, an AEM flow cell is assembled with the NW1M1R electrocatalyst, which serves as a cathode, and the homemade NiFe LDH serves as the anode (Fig. S28†). The polarization curves of the AEM are shown in Fig. 2f; commercial Pt/C is added for comparison. Notably, the NiFe LDH (+)‖NW1M1R (−) couple achieves a current density as high as 1 A cm−2 at 2 V, which is superior to that of the commercial NiFe LDH (+)‖Pt/C (−) and most of the reported electrocatalyst (Fig. S29†). More importantly, the stability of the assembled AEM flow cell is investigated at a constant current density of 1 A cm−2, which exhibits excellent durability for 90 hours, indicating the NW1M1R electrode to be a promising HER electrocatalyst in the AEM water electrolyzer (Fig. 2g).
To reveal the effect of each component on the performance enhancement, we first employed density functional theory (DFT) to gain insights into the H* absorption/desorption trend. As shown in Fig. S30,† the ΔGH of Ni17W3, WO3, and MoO3 is −0.60, 1.57, and 0.86 eV, respectively. Thus, the NW1M1R constructed Ni17W3 with a negative value and WO3/MoO3 with a positive value, indicating the potential hydrogen spillover process between the metal and metal oxide interface.34,35 Then, the in situ Raman technique was utilized to monitor the phase evolution during the HER process. For the NWP sample, it is clear that the full width at half maximum (FWHM) of the O–W–O vibrational peak located at 325 cm−1 increases as the reaction proceeds, indicating a decrease in the crystallinity of WOx (Fig. 3a).36 Similarly, the characteristic peaks of W–O–W located at 820 and 843 cm−1 decreased when the negative voltage was applied.37 The spontaneous disappearance in an alkaline environment contributes to surface reconstruction. Importantly, a new Raman signal appeared at 1055 cm−1 when the voltage reaches −0.35 V,38 which belongs to the metal oxide hydration vibration. In this way, metal oxide combines with water to form metal hydroxides, which further transform into hydrogen bronze with the assistance of electrons (eqn (1) and (2)). It should be noted that the hydrogen atoms in hydrogen bronze will dissociate into highly mobile protons and electrons. The protons are localized near the O2− in the lattice, leading to hydroxide character, whereas the electrons localize on the metal d-orbitals, reducing the M6+ to M5+. More importantly, the saturated concentration of protons will result in H2 generation on the metal oxide surface (eqn (3)).
MVIO3 + H2O ↔ M(OH)x | (1) |
M(OH)x + H2O + e− ↔ HyMVMVIO3−z | (2) |
HyMVMVIO3−z ↔ y/2H2↑ + MVIO3 + e− | (3) |
For the NW1M1P sample, the characteristic peak of Mo–O stretching vibration appears at 887 cm−1 (Fig. 3b).37 Meanwhile, the location of the characteristic peak of WO shows a slight redshift, indicating that MoO3 induces a change in the lattice stress–strain behavior.39 Unlike the NWP sample, the W–O–W signal located at 820 and 843 cm−1 consistently has distinct characteristic peaks in the low voltage range (<−0.3 V), suggesting a stronger interaction between MoO3 and WO3, which is in agreement with the XPS results. Interestingly, the characteristic peaks of metal oxide hydration appear at a low voltage of −0.2 V, completing the phase transition from the metal oxide to the real active phase hydrogen bronze more quickly. The lower activation potential of the NW1M1R electrocatalyst is supposed to be due to the fact that the high oxygen vacancies concentration is beneficial for the hydrogen insertion/extraction process.
In addition, due to the different HER kinetics based on the Tafel slope analysis, the in situ EIS technique was used to characterize the detailed mass/charge transfer kinetic behavior. Both samples were subjected to EIS analysis at different voltages (Fig. 3c and d). Bode plots can be divided into two regions, high-frequency and low-frequency, and the signals in each of these regions can be used to analyze the conductivity of the catalysts at the inner layer and the interface as well as the charge transfer between the catalysts and the electrolyte.40 It is clear that the kinetics of NW1M1R and NWP are significantly different. NW1M1R has characteristic peaks only at low frequencies, indicating that the kinetic process is restricted to the charge/mass transfer process at the electrode/electrolyte interface.41 Meanwhile, the almost negligible inner-layer/interface resistance can be attributed to the unique morphology and the SMSI effect. However, for NWP, the lower phase angle in the low-frequency area and the higher phase angle in the high-frequency area demonstrated that the kinetic process is mainly controlled by the inner layer/interface charge transfer resistance.33 Then, the double-parallel equivalent circuit model is established based on the recorded Nyquist plots (Fig. S31 and S32†). The time constants of CPE1 and R1 can represent the hydrogen adsorption behavior, while CPE2 and R2 are related to the mass/charge transfer kinetics (Tables S2 and S3†).14 Firstly, the lower resistance (R1) of NW1M1R across the entire voltage range suggests efficient mass/charge transfer kinetics, in line with Tafel slope analysis. Then, the detailed H* adsorption/desorption behavior was further investigated. The H* adsorption charge can be represented by integrating the CPE2-T with the overpotential (Fig. 3e). The larger area of NW1M1R than that of NWP indicates that NW1M1R adsorbs a larger amount of H*, and the higher H* coverage on the surface is conducive to the Volmer–Tafel process. In addition, the H* adsorption behavior of these two samples exhibits the potential-dependent increase as soon as the cathodic potential is applied, indicating that the hydrogen spillover phenomenon occurs at the interface; only the transfer efficiency from metal to support is different. Then, the desorption process of H* is investigated by analyzing the EIS-derived Tafel slope (Fig. 3f). The smaller slope proves that NW1M1R has a larger H* desorption charge, demonstrating that NW1M1R simultaneously has faster hydrogen desorption kinetics. In conclusion, compared to NWP, the kinetics of NW1M1R are limited only by the H* charging/mass transfer process and exhibit superior H* adsorption/desorption kinetics.
Ni17W3 + H2O + e− ↔ Ni17W3–H* + OH− | (4) |
Ni17W3–H* + Ni17W3–H* + e− ↔ Ni17W3 + H2 | (5) |
Ni17W3–H* + MO3 ↔ Ni17W3 + HyMVMVIO3−z | (6) |
Ni17W3–H* + H2O + e− ↔ Ni17W3 + H2↑ + OH− | (7) |
Based on the analysis results, the HER process is detailed below. For the NM1W1R sample, the obvious hydrogen spillover occurs at the metal/metal oxide interface. First, the adjacent H* combine with each other on the metal surface to produce H2 (eqn (4) and (5)). Then, due to the abundant oxygen vacancies of metal oxide support, the H* insertion/extraction proceeds easily in the metal oxide, resulting in a faster H* transfer from the metal to the metal oxide, leading to the metal oxide support surface, also producing significant amounts of hydrogen (eqn (3) and (6)). After increasing cathodic potential, the hydration capacity of metal oxides increased significantly, further promoting the phase transition process from metal oxides to hydrogen bronzes (eqn (2) and (3)). Thus, in turn, the whole H* adsorption, transfer, and desorption kinetic increased on the catalyst surface and interface, resulting in a superior HER performance. However, in the case of NMP, the lower H* coverage on the surface suggests that the transfer of H* from the metal to the metal oxide support is inefficient. The H2 is produced primarily from the combination of H* on the surface and H2O in the electrolyte, resulting in sluggish H* adsorption/desorption kinetics (eqn (4) and (7)). Meanwhile, the noticeable hydration of the metal oxide support requires a larger potential drive, and the kinetic behaviors of both hydrogen spillover and hydrogenation are relatively slow, resulting in unsatisfactory catalytic activity (Fig. 3g and h).
In order to verify whether the hydrogen spillover and metal oxide hydrogenation work at a wider pH range, the electrochemical performance of NM1W1R is investigated in 0.5 M H2SO4 (pH = 0.21) and 1.0 M PBS (pH = 7.2) electrolyte, respectively. Surprisingly, the same trend was observed under acidic and neutral conditions as under alkaline conditions. As shown in Fig. 4a and b, NW1M1R has the best HER performance (14 mV in 0.5 M H2SO4 and 42 mV in 1.0 M PBS) compared to NWP (61 mV in 0.5 M H2SO4 and 86 mV in 1.0 M PBS), NW1M1.5R (44 mV in 0.5 M H2SO4 and 50 mV in 1.0 M PBS), NW1M0.5R (42 mV in 0.5 M H2SO4 and 59 mV in 1.0 M PBS) and Pt/C (39 mV in 0.5 M H2SO4 and 46 mV in 1.0 M PBS). The Tafel slope corresponding to NW1M1R (32.6 mV dec−1 in 0.5 M H2SO4 and 73.9 mV dec−1 in 1.0 M PBS) is similarly lower than that of NWP (66.6 mV dec−1 in 0.5 M H2SO4 and 76.3 mV dec−1 in 1.0 M PBS), NW1M1.5R (41.9 dec−1 in 0.5 M H2SO4 and 77.7 mV dec−1 in 1.0 M PBS), NW1M0.5R (51.1 mV dec−1 in 0.5 M H2SO4 and 77.3 mV dec−1 in 1.0 M PBS) and Pt/C (37.03 mV dec−1 in 0.5 M H2SO4 and 81.1 mV dec−1 in 1.0 M PBS) (Fig. 4c and d). The long-term HER performance of NW1M1R at different current densities and in different electrolytes (300 h in 0.5 M H2SO4 and 140 h in 1.0 M PBS) shows no significant upward trend in voltage (Fig. 4d and e). These results demonstrate that NW1M1R could serve as a superior pH-universal electrocatalyst for HER.
Footnote |
† Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d4ta00729h |
This journal is © The Royal Society of Chemistry 2024 |