Xijie
Chen
a,
Fengming
Zhang
a,
Xiao
Wang
a,
Fangming
Liu
a,
Jinhan
Li
a,
Meng
Yu
*a and
Fangyi
Cheng
*ab
aKey Laboratory of Advanced Energy Materials Chemistry (Ministry of Education), Engineering Research Center of High-efficiency Energy Storage (Ministry of Education), College of Chemistry, Nankai University, Tianjin 300071, China. E-mail: nkyu2023@nankai.edu.cn; fycheng@nankai.edu.cn
bHaihe Laboratory of Sustainable Chemical Transformations, Tianjin 300192, China
First published on 26th December 2024
The alkaline hydrogen evolution reaction (HER) is plagued by intricate interfacial reactions involving the dissociation of interfacial H2O molecules and adsorption/desorption of Hads/OHads species, which impede the practical application of water electrolysis. Herein, a self-supported Co9S8@CoMoPx electrode with a nanosheet cluster morphology was developed using a stepwise electrodeposition method for an efficient electrocatalytic HER. Benefiting from the coexistence of multivalent metal sites, the Co9S8@CoMoPx electrode integrated both oxophilic and protophilic properties to facilitate the cracking of molecular H2O and subsequent hydrogen generation. As a result, the obtained Co9S8@CoMoPx electrode exhibited superior alkaline HER activities, delivering an overpotential of 226 mV at −500 mA cm−2 with a low attenuation rate of 11 μV h−1 after 1000 h. An anion-exchange membrane water electrolysis device was then assembled by matching the Co9S8@CoMoPx cathode with an NiFe-based anode to demonstrate its industrial application potential. This work emphasizes the significance of constructing multivalent metal sites to simultaneously achieve oxophilicity and protophilicity, providing a guideline for the rational design of heterostructure electrocatalysts for efficient energy conversion.
Broader contextSustainable anion-exchange membrane water electrolysis (AEMWE) is recognized as a promising candidate to mitigate the current carbon emission issues and realize the future hydrogen economy. However, as the cathodic reaction, the hydrogen evolution reaction (HER) is plagued by the dissociation of interfacial H2O molecules and the adsorption/desorption of Hads/OHads species, which result in sluggish water dissociation and high energy loss in green hydrogen production. In this work, we report a double-layered Co9S8@CoMoPx self-supported electrode with multivalent active sites to address intricate interfacial reactions in the alkaline HER. Multivalent metal sites on the surface endow the Co9S8@CoMoPx electrocatalyst with both low-valence protophilic sites and high-valence oxophilic sites, thus accelerating catalytic kinetics and delivering low overpotentials of 41 and 226 mV at current densities of −10 and −500 mA cm−2. Meanwhile, Co(OH)x on the surface can impede the leaching of MoO42− ions through electrostatic interaction and enhance catalytic stability for long-term operation (1000 h at −500 mA cm−2). This work not only emphasizes the significance of constructing composite electrocatalysts with multivalent metal sites to simultaneously achieve oxophilicity and protophilicity but also provides a reference for the precise customization of self-supported heterostructure electrocatalysts for efficient energy conversion. |
To overcome this kinetic barrier, interface engineering strategies conducted on heterogeneous catalysts have been critically explored.15,16 For example, Pt nanoparticles were loaded onto MgO nanosheets with oxygen vacancies, whereby MgO could promote H2O dissociation and H3O+ accumulation around Ptδ− sites to form an acid-like local microenvironment, which accelerated the kinetics of the alkaline HER.17 Moreover, CeO2 and CoS2 materials were coupled to form a heterostructure, in which CeO2 provided strong acidic Ce sites to promote the adsorption and dissociation of water, while CoS2 offered weakly basic S sites for the surface transfer of Hads and subsequent H2 evolution.18 Although the above heterostructure electrocatalysts delivered impressive alkaline HER performances, their synergistic effect was trapped at the limited interface between the different components. For overcoming this, constructing electrocatalysts with mixed multivalent metal ions would be beneficial for integrating multifunctional reaction sites with proper H2O dissociation or Hads/OHads adsorption ability at the atomic level. Besides, the combination of a surface protective layer and internal active material with a partially leachable component can improve the stability without having a significant impact on catalytic activity.19,20 Therefore, a heterostructure electrocatalyst with the coexistence of a protective layer and multivalent active sites could be expected to drive stable cathodic HER processes in alkaline conditions.
In this research, a double-layered Co9S8@CoMoPx self-supported cathode with multivalent active sites was designed and fabricated by a continuous cathodic layer-by-layer assembly method. Owing to the introduction of the Mo element, the inherent electronic structure of Co9S8@CoMoPx was modulated, which contributed to the existence of multivalent Co metal sites on the surface at the atomic level and endowed the Co9S8@CoMoPx electrode with both low-valence protophilic sites and high-valence oxophilic sites. Detailed electrochemical investigations were performed and demonstrated that Co9S8@CoMoPx delivered excellent catalytic performance with a low overpotential of 226 mV to reach −500 mA cm−2 and long-term stability for 1000 h by inhibiting MoO42− leaching through an electrostatic anchoring effect of the high-valence sites in the Co9S8 layer. To explore the prospect of practical industrial applications, the Co9S8@CoMoPx cathode was also scaled-up and assembled into an AEMWE device coupled with a Ni3S2@NiFePx anode. The results in the present study highlight the significance of multivalent active sites and a protective layer for realizing an efficient and stable alkaline HER process.
![]() | ||
Fig. 1 (a) Schematic of the synthesis route for the Co9S8@CoMoPx electrode. CoMoPx stands for amorphous cobalt–molybdenum-based phosphide. Surface SEM images of (b) CoMoPx, (c) Co9S8 and (d) Co9S8@CoMoPx electrodes. (e) Cross-sectional SEM image of Co9S8@CoMoPx and corresponding (f) elemental line scanning following the arrow in Fig. 1(e). Related high-resolution transmission electron microscopy (HRTEM) images for (g) CoMoPx and (h) Co9S8. |
In addition, the optimized deposition times of the bottom and surface layer were further investigated by considering the electrochemical performance. As depicted in Fig. S7a (ESI†), when the deposition time of the CoMoPx layer was 15 min, the double-layered electrode showed the best HER performance. For the top Co9S8 layer, 3 min was determined to be an appropriate deposition time to ensure full coverage with a negligible effect on the electrocatalytic activity (Fig. S7b, ESI†). The cross-sectional SEM image of the Co9S8@CoMoPx electrode, prepared by 15 min and 3 min depositions of the CoMoPx bottom layer and Co9S8 surface layer, verified the two-layered structure of the composite (Fig. 1e). Considering the difficulty in distinguishing the signals of Mo and S in the SEM element mapping due to the closely located characteristic peaks of Mo Lα (2.29 keV) and S Kα (2.3 keV), the signal for P was used to identify the bottom and surface layers. According to Fig. S8 (ESI†), the P element was mainly distributed at the bottom, confirming the inner and outer layers were CoMoPx and Co9S8. Besides, element line scanning following the direction of the arrow in Fig. 1e was conducted (Fig. 1f). It was found that the signal for P gradually increased after about 0.65 μm, implying that the thicknesses of the surface Co9S8 and bottom CoMoPx were ∼0.65 μm and ∼0.6 μm, respectively.
Transmission electron microscopy (TEM) images and selected area electron diffraction (SAED) patterns were obtained and confirmed the amorphous and polycrystalline structures of CoMoPx (Fig. S9, ESI†) and Co9S8 (Fig. S10, ESI†), respectively. The related high-resolution transmission electron microscopy (HRTEM) images in Fig. 1g showed no evident crystal structure for CoMoPx and obvious lattice fringe spacings of 0.29 and 0.25 nm, corresponding to the (222) and (400) planes of Co9S8.21,22 Moreover, the phase of Co9S8 was also examined by X-ray diffraction (XRD) (Fig. S11, ESI†), which gave consistent results with the TEM results.23,24 Inductively coupled plasma-optical emission spectrometry (ICP-OES) was performed and the results showed that the Co/Mo ratio in the Co9S8@CoMoPx double-layered composite was about 5:
1 (Table S1, ESI†).
According to the polarization curves in Fig. 2a and Fig. S12, S13, ESI,† Co9S8@CoMoPx demonstrated superior HER performance (41/226 mV@−10/500 mA cm−2) than CoP (188/374 mV@−10/500 mA cm−2), Co9S8 (445 mV@−10 mA cm−2), and almost similar to the single-layer CoMoPx electrode (38/212 mV@−10/500 mA cm−2). The corresponding Tafel slope values of Co9S8@CoMoPx and CoMoPx were 82.3 and 78.7mV dec−1, which were lower than those for CoP (87.0 mV dec−1) and Co9S8 (107.1 mV dec−1) without Mo (Fig. 2b), manifesting the accelerated HER kinetics.25–27 Electrochemical impedance spectroscopy (EIS) measurements were performed for Co9S8@CoMoPx at −0.1 V vs. RHE and showed there was a rapid charge-transfer rate at the electrode–electrolyte interface (Fig. 2c). Meanwhile, operando EIS was also carried out to determine the electron-transfer resistance and reaction kinetics during the HER process. As displayed in Fig. S14 (ESI†), the Rct values of Co9S8@CoMoPx and CoMoPx were obviously smaller than those of CoP and Co9S8 under the same overpotential, illustrating that greater charges were involved in the Faraday reaction on the Co9S8@CoMoPx and CoMoPx surfaces instead of being stored at the interface.28,29 Similarly, the phase angles of Co9S8@CoMoPx and CoMoPx decreased more rapidly (Fig. S15, ESI†), further confirming the accelerated charge-transfer process.30,31 The Cdl value of the Co9S8@CoMoPx electrode was higher than that of other single-layer electrodes (Fig. S16, ESI† and Fig. 2d), implying the double-layered composite fabricated by the two-step layer-by-layer cathodic deposition endowed the electrode with a rougher surface and more active sites.
![]() | ||
Fig. 2 (a) LSV curves of different electrodes in 1.0 mol L−1 KOH with 95% iR-compensation and (b) the corresponding Tafel slopes. (c) Nyquist plots of electrocatalysts measured at −0.1 V (vs. RHE) and (d) double-layer capacitances. (e) Accelerated duration test of Co9S8@CoMoPx at a 50 mV s−1 scan rate. (f) Chronopotentiometry curves of Co9S8@CoMoPx and CoMoPx at −500 mA cm−2. (g) Inductively coupled plasma-optical emission spectrometry (ICP-OES) results of the electrolyte after the HER chronopotentiometry tests. The detailed data are listed in Table S2 (ESI†). (h) HER performance and (i) stability comparison of Co9S8@CoMoPx with CoMoP-based catalysts and some other HER catalysts reported in the literature, with the detailed information summarized in Table S3 (ESI†). |
To further evaluate the catalytic stability, accelerated duration tests and chronopotentiometry measurements were performed. As presented in Fig. 2e, the polarization curve of Co9S8@CoMoPx after 5000 cyclic voltammetry (CV) cycles almost overlapped with the original linear sweep voltammetry (LSV) curve. The attenuation of Co9S8@CoMoPx during the accelerated duration test at −100 mA cm−2 was only 4 mV, while that for the CoMoPx electrode was 30 mV (Fig. S17, ESI†), verifying the superior cycling stability of Co9S8@CoMoPx. According to the chronopotentiometry test results (Fig. 2f), although the initial HER performance of CoMoPx was slightly better than that of Co9S8@CoMoPx, the catalytic activity seriously declined during continuous operation. In contrast, Co9S8@CoMoPx showed impressive operation stability at −500 mA cm−2 for 1000 h with a low attenuation rate of 11 μV h−1. In addition, ICP-OES was applied to analyse the origin of the difference in stability between Co9S8@CoMoPx and CoMoPx by measuring the dissolved metal contents in the electrolyte after hydrogen evolution (Fig. 2g). A lower Mo content was observed in the electrolyte for Co9S8@CoMoPx after the HER than with the single-layer CoMoPx. The above experiments firmly prove that the polycrystalline Co9S8 layer prolonged the operation life of the electrocatalyst by inhibiting the dissolution of Mo sites in the inner CoMoPx layer. Besides, the microstructure morphology of the Co9S8@CoMoPx electrode could be well-retained with robust features under abundant hydrogen gas evolution (Fig. S18, ESI†). Compared with some other recently reported CoMoPx-based electrocatalysts, the Co9S8@CoMoPx electrode with adjacent oxophilic and protophilic sites demonstrated both impressive HER activity and stability, especially under a larger current density, as shown in Fig. 2h and i.
Next, X-ray photoelectron spectroscopy (XPS) was carried out to probe the chemical composition and elemental valence state of the electrode before and after reaction. The Co 2p spectrum of CoMoPx displayed signals for Co0 (∼777.8 and ∼792.8 eV), Co3+ (∼781.0 and ∼796.5 eV), Co2+ (∼782.5 and ∼798.0 eV), and satellite peaks (∼786.2 and ∼802.8 eV) (Fig. S19a, ESI†).32,33 The Mo 3d spectrum was divided into four components: Mo0 (∼227.4 and ∼230.5 eV), Mo4+ (∼228.5 and ∼231.6 eV), Mo5+ (∼230.3 and ∼233.4 eV), and Mo6+ (∼232.1 and ∼235.2 eV) (Fig. S19b, ESI†).32,34 The P 2p profile presented two peaks at 129.1 and 133.2 eV (Fig. S19c, ESI†), attributed to the characteristic peaks of M–P (M stands for transition metal) and P–O signals, respectively.35 The Co 2p pattern of Co9S8 illustrated the predominant +2 and +3 valence states for the Co element in Co9S8 (Fig. S20a, ESI†). The non-metallic S 2p spectrum presented two peaks at ∼163.0 and ∼168.6 eV, which were attributed to M–S and S–O peaks, respectively (Fig. S20b, ESI†).36 As for Co9S8@CoMoPx, according to the Co 2p spectrum (Fig. 3a) of Co9S8@CoMoPx before the chronopotentiometry test, characteristic peaks at 796.5 eV (Co 2p1/2) and 780.7 eV (Co 2p3/2), along with two satellite peaks at 803.0 and 785.8 eV could be observed, indicating the presence of Co2+ and Co3+. After the chronopotentiometry test, the Co element in the Co9S8@CoMoPx composite became more multivalent with the coexistence of 0, +2, and +3 chemical states (Fig. 3a). While the Mo element mainly existed in a +6 valence state (MoO42−) on the surface after reaction (Fig. 3b).34 Meanwhile, no significant peak was observed in the P 2p diagram after the HER (Fig. 3c) due to the element leaching, as the thick outer layer of Co9S8 may block the XPS signal of the underlying CoMoPx layer. To further rigorously support this speculation, we collected XPS depth profiling of the reacted Co9S8@CoMoPx using the Ar+ ion, which verified that the P content increased with the etching thickness (Fig. S21, ESI†). Moreover, the content of the surface S element decreased and M–S was the dominant state after the HER (Fig. 3d). To further investigate the reason for the coexistence of multiple Co valence states, the XPS spectra of the electrodes with (CoMoPx) and without Mo (Co9S8 and CoP) after hydrogen evolution were also collected. After the HER reaction, the Co 2p spectrum of CoMoPx also displayed the feature with multivalent Co containing 0, +2, and +3 valence states (Fig. S19d, ESI†), while the Mo 2p spectrum showed signals for Mo0 and Mo6+ (Fig. S19e, ESI†), in which Mo6+ was the dominant state (MoO42−). In contrast, the Co 2p spectrum of Co9S8 after the HER contained only the signal peaks for Co2+ and Co3+ (Fig. S20c, ESI†), while the S 2p spectrum could be divided into M–S and S–O peaks (Fig. S20d, ESI†). Besides, the Co 2p spectrum for CoP exhibited signal peaks for Co0, Co2+, and Co3+ (Fig. S22a, ESI†). Compared with the Co 2p spectrum of CoP before the HER, CoMoPx with the Mo component possessed a higher proportion of Co0, manifesting a possible electronic regulation effect between the Co and Mo atoms. Moreover, the Co ions with multivalent states in the Co 2p spectrum for CoP before the HER turned into a narrower valence distribution of Co2+ and Co3+ after the reaction (Fig. S22c, ESI†). The P 2p spectra of CoP demonstrated a decrease in P element near the surface (Fig. S22d, ESI†).
The valence states distribution of Co and Mo for the obtained electrocatalysts before and after HER progress was summarized in Fig. 3e and f. The Co element in Co9S8@CoMoPx and CoMoPx containing Mo components showed the coexistence of multivalent states, including 0, +2, and +3, while that of Co9S8 and CoP without the Mo element predominantly showed the +2 and +3 valence states after the HER. As the Pourbaix diagram in aqueous solution depicts, the Mo element mainly existed as Mo6+ (MoO42−) on the surface of Co9S8@CoMoPx due to the reduction potential of MoO42− being lower than the hydrogen evolution potential in alkaline conditions (Fig. 3f).37 At this point, the coexistence of multivalent states of the Co element could be attributed to two points: on the one hand, abundant OH− diffusing from the bulk solution or generated by interfacial H2O dissociation will adhere to the catalyst surface, and then promote the surface hydroxylation to generate high-valent Co2+/3+;38 on the other hand, introducing Mo into the electrocatalyst influences the initial electronic structure. Owing to the existence of the element with high electronegativity, Mo provides more electrons instead of partial Co to P ions, which is beneficial for maintaining the low-valent Co0 and then broadening the valence state range. The Raman spectra of Co9S8@CoMoPx before and after HER were recorded and are shown in Fig. 3g. Typical bands of Co9S8 at 468, 515, and 658 cm−1 were still retained.39 In addition, characteristic peaks of MoO42− (894 cm−1)37,40–43 and partially oxidized Co(OH)x (503 and 611 cm−1) appeared after the reaction.4,43 These further elucidate the coexistence of multivalent metal sites on the surface of Co9S8@CoMoPx. Based on the above discussion, the P and S non-metallic elements leach into the electrolyte during the HER process, while OH− in the alkaline solution attaches on the electrode and affects the valence distribution of metal elements on the surface. Thus, the multivalent states of metal sites with Co0, Co2+/3+(Co(OH)x), and Mo6+(MoO42−) coexisted on the electrode surface. The H* in the H2O molecule can be adsorbed at the low-valent Co0 sites and OH* is adsorbed at the oxophilic Mo6+ metal sites to break the H–OH bond and allow the Volmer step to proceed.40,44 The H* is subsequently adsorbed at the metal sites with a low-valent state to accomplish the Heyrovsky reaction. In parallel, the presence of Co(OH)x on the surface can anchor MoO42− anions,43 thereby inhibiting the leaching of active sites and improving the catalytic stability of the Co9S8@CoMoPx (Fig. 3h). In conclusion, the multivalent Co element in Co9S8@CoMoPx provides both protophilic low-valence states and oxophilic high-valence states, which is beneficial to accelerate the alkaline HER kinetics with intricate interfacial reactions.
To explore the wettability of the Co9S8@CoMoPx electrode, contact-angle (CA) measurements were then performed, as shown in Fig. 4a and b. The static water contact-angles of Co9S8@CoMoPx and bare NF were 0° and 111.1°, while the bubble contact-angles were 139.6° and 92.6°, respectively. The nearly zero droplet contact-angle and larger bubble contact-angle manifested the hydrophilic property of the Co9S8@CoMoPx surface. This property allows the electrolyte to flow more evenly through the Co9S8@CoMoPx electrode and reduces the “dead zone” during the generation of plentiful hydrogen bubbles.45,46 Furthermore, the size of the Co9S8@CoMoPx electrode was amplified from 1.0 × 1.0 to 10.0 × 10.0 cm2 (Fig. 4c), confirming the potential for industrial application of Co9S8@CoMoPx as a cathode.
![]() | ||
Fig. 4 (a) Static water and bubble contact-angle measurements of the Co9S8@CoMoPx electrocatalyst and blank NF. (b) Comparison diagram of the contact-angle values in Fig. 4a. (c) Scaled-up image of the Co9S8@CoMoPx electrode. |
The self-supported Co9S8@CoMoPx cathode coupled with a Ni3S2@NiFePx anode was assembled in an alkaline exchange membrane water electrolysis (AEMWE) device to validate its prospects for industrial application (Fig. S23, ESI†).19 The liquid feeding mode and operating temperature of the device were optimized. According to the polarization curves in Fig. 5a and b, the catalytic activity with a 1.0 mol L−1 KOH feed from anodic side delivered the most prominent enhancement because of the increased OH− concentration at the membrane electrode interface, which accelerated the reaction kinetics and benefitted the release of high-purity hydrogen gas. Moreover, the effect of operation temperature was also explored. As shown in Fig. 5c, the performance of the AEMWE device was enhanced with increasing temperature, as a higher temperature could effectively reduce the Rct of the electrolyser through accelerating the kinetics of charge transfer and the interfacial electrochemical reaction (Fig. 5d). The Co9S8@CoMoPx‖Ni3S2@NiFePx-based AEMWE device delivered a current density of 1 A cm−2 by applying 1.96 V at 80 °C with the anode feed. The whole electrolyser could stably operate for 30 h under a constant current density of 500 mA cm−2 (Fig. 5e) and could be maintained for 10 h during an accelerated attenuation test (Fig. 5f).
![]() | ||
Fig. 5 (a) Polarization curves and (b) Nyquist plots of AEMWE with the Co9S8@CoMoPx cathode at 25 °C in 1.0 mol L−1 KOH. (c) Polarization curves and (d) Nyquist plots of Co9S8@CoMoPx-based AEMWE with anode feed at different temperatures. (e) Chronopotentiometry curves at 500 mA cm−2 and (f) accelerated degradation test of the Co9S8@CoMoPx‖Ni3S2@NiFePx-based AEMWE device. An optical picture of the AEMWE device is shown in Fig. S23 (ESI†), and the area sizes of the Co9S8@CoMoPx and Ni3S2@NiFePx electrodes were both 2 × 2 cm2. |
Footnote |
† Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d4ey00252k |
This journal is © The Royal Society of Chemistry 2025 |