Pengliang
Sun‡
ac,
Eduardo
Gracia-Espino‡
b,
Fang
Tan‡
ac,
Hua
Zhang
ac,
Qingquan
Kong
d,
Guangzhi
Hu
*ac and
Thomas
Wågberg
*be
aDonghai Laboratory, Zhoushan 316021, China. E-mail: guangzhihu@ynu.edu.cn
bDepartment of Physics, Umeå University, Umeå S-90187, Sweden. E-mail: thomas.wagberg@umu.se
cInstitute for Ecological Research and Pollution Control of Plateau Lakes, School of Ecology and Environmental Science, Yunnan University, Kunming 650504, China
dSchool of Mechanical Engineering, Chengdu University, Chengdu 610106, China
eWallenberg Initiative Material Science for Sustainability, Department of Physics, Umeå University, Umeå S-901 87, Sweden
First published on 22nd May 2024
Electrochemical water splitting powered by renewable energy sources hold potential for clean hydrogen production. However, there is still persistent challenges such as low solar-to-hydrogen conversion efficiency and sluggish oxygen evolution reactions. Here, we address the poor kinetics by studying and strengthening the coupling between Ce and W, and concurrently establishing Ce–W bi-atomic clusters on P,N-doped carbon (WN/WC–CeO2−x@PNC) with a “treasure-bowl” style. The bifunctional active sites are established using a novel and effective self-sacrificial strategy involving in situ induced defect formation. In addition, by altering the coupling of the W(d)–N(p) and W(d)–Ce(f) orbitals in the WN/WC–CeO2−x supramolecular clusters, we are able to disrupt the linear relationship between the binding energies of reaction intermediates, a key to obtain high catalytic performance for transition metals. Through the confinement of the WN/WC–CeO2−x composite hetero-clusters within the sub-nanometre spaces of hollow nano-bowl-shaped carbon reactors, a stable and efficient hydrogen production via water electrolysis could be achieved. When assembled together with a solar GaAs triple junction solar cell, a solar-to-hydrogen conversion efficiency of 18.92% in alkaline media could be realized. We show that the key to establish noble metal free catalysts with high efficiency lies in the fine-tuning of the metal–metal interface, forming regions with near optimal adsorption energies for the reaction intermediates participating in water electrolysis.
New conceptsSingle-cluster site catalysts with uniform metal active sites have tremendous potential in classical catalytic reactions. However, research on the introduction of oxygen vacancy defects into hetero-clusters to enhance the electronic structure for solar-driven water splitting remains unreported. Here, we propose a novel and effective self-sacrificial in situ induced defect strategy using phosphotungstic acid supramolecular clusters. This strategy demonstrates excellent electrochemical performance and solar-to-hydrogen conversion efficiency. The synergistic effect at the Ce–W bi-atomic clusters not only lowers the adsorption energy of reaction intermediates but also facilitates electronic rearrangement, effectively tuning local conductivity, and charge density, thereby enhancing catalytic activity. The single-cluster site catalyst unveils a path for renewable fuel production towards a carbon-free green energy generation. |
Among these, tungsten-based electrocatalysts are strong candidates to catalyse the OER and HER because of their abundancy, their environmentally benign properties, and high intrinsic catalytic activity.12 It has been reported that transition metal carbides (TMCs) (TaC,13 Mo2C,9 and WC14) and transition metal nitrides (TMN) (MoN,15 Mo2N3,16 and W2N12) have attractive attributes with metal-like physicochemical properties and well-adapted electronic structures. Here, we note that the d-band electronic structure of WC14 is similar to that of Pt near the Fermi levels (EF), which potentially provides for a swift Hads release, one of the descriptive parameters of an efficient HER. Sun et al.17 reported a catalytic electrode based on a N-doped WC nanoarray. They could show that N-doping optimises the hydrogen bonding by adjusting the surface energy levels, thus kinetically accelerating the reaction. However, this effect is limited to acidic systems and the underlying mechanism remains unclear. Other studies taking advantage of the similarity of W-based nitrides and precious metals (Pd and Pt) when interacting with hydrogen atoms, was reported by Fu et al.18 who synthesized WN on RGO using H3[PO4(W3O9)4] anion clusters as the W source. They showed that the interaction of P with RGO and WN increased the work function, close to that of Pt metal. The surface work function has been shown to correlate well with the metal–hydrogen bond strength16,19 and the hydrogen adsorption energy weakens linearly with an increased work function.20–22 It is thus clear that doping or hybridisation can optimise the W–Had bond to promote the adsorption of hydrogen and desorption of Had, further improving the HER performance,.23 However, whether water dissociation or OHad desorption can be promoted on the surface of tungsten nitride in alkaline HER remains unclear. An interesting strategy to facilitate both adsorption/dissociation and hydrogen adsorption is to establish heterostructures where the two different surfaces can form heterojunctions that can accelerate the hydrogen diffusion step by and promote the subsequent Heyrovsky step.24
Cluster catalysts represent an interesting form of materials where the redox states can be affected compared to their bulk surface counterparts,25 and thereby optimise the adsorption/dissociation of water and the formation of H*/OH* intermediates. They can also serve as model systems making it possible to get insight into structure/compositional effects on water splitting mechanisms. However, so far most metal nitrides and carbides have been synthesized as nanoparticles with no corresponding cluster structure. Constructing fully exposed nanoclusters remains however challenging because in a reaction system, the metal/non-metallic elements usually form separate phases.
Herein, we utilize a simultaneous carbonisation of different metals in a limited space to form a variety of hetero-element dopants, resulting in a three-dimensional (3D) porous electrocatalyst comprising nitrogen- and phosphorus-doped bowl-shaped carbon encapsulating WN/WC–CeO2−x hetero-clusters. We find that, when used for water splitting reactions in alkaline media, the 3D open porous structure of WN/WC–CeO2−x allows for a good mass transfer and gas evolution while maintaining a stable solid–liquid interface. High-density heterogeneous metal clusters provide an abundance of active sites, while the bowl structure facilitates the diffusion of reactants. Moreover, the interconnected bowl-shaped carbon improves the conductivity and structural stability of the encapsulated nanoclusters during electrocatalysis. DFT calculations reveal that oxygen vacancies in cerium oxide weaken the adsorbed interaction between metal species and enable f–d orbital coupling through the entry of the Ce atoms into the surface of W, which results in suitable binding energies for the H-intermediates to efficiently generate H2. A solar-driven water electrolyser employing the electrocatalyst achieves a stable, sustainable green hydrogen production with a solar-to-hydrogen (STH) conversion efficiency of 18.92%. Our study presents a viable strategy for designing fully exposed nanocluster catalysts for advancing the field of electrocatalytic water splitting.
The projected density of states (PDOS) (Fig. 1(d) and (e)) was computed to analyse the electronic properties of WN/WC–CeO2−x. Compared to WN/WC and WN/WC–CeO2 (Fig. S3, ESI†), the introduction of oxygen vacancies, (Vo) and Ce induced a pronounced distribution of W states near the Fermi level in WN/WC–CeO2−x, accompanied by a higher density of metallic states, indicating enhanced electron transfer at the interface of WN/WC–CeO2−x. The d-band centre (εd) is an important indicator of the bonding stability and strength of catalytic intermediates.26 The broader d-band and higher εd in proximity to the Fermi level in WN/WC–CeO2−x (Fig. S4, ESI†) indicate increased occupancy of the antibonding orbitals, which favours the adsorption of reaction intermediates, consequently accelerating the reaction. The εd of WN/WC–CeO2 and WN/WC–CeO2−x maintained distinct offsets relative to Ef, with values of –2.339 and –2.357 eV, respectively (Fig. S4, ESI†). The oxygen vacancies induce disparities between the d-band centres of Ce and W active sites, consequently lowering the free energy of the rate-determining step (RDS). Coupling WN/WC with Vo-defective CeO2−x created additional active sites and enhanced the conductivity, facilitating efficient charge transfer during catalysis. The coupling also modulates the electronic structure of the Ce–W species, which lowers the formation energy barrier for reaction intermediates (Fig. 1(f)). This fine-tuning of the electronic structure enhances the intrinsic activity of the resulting electrode and further improves its catalytic performance.
The surface morphology of WN/WC–CeO2−x@PNC was characterized using field emission-scanning electron microscope (FE-SEM). The WN/WC–CeO2−x hetero-clusters are uniformly distributed on the bowl-shaped carbon (Fig. 2(b)); and SEM of the samples during synthesis (Fig. S5, ESI†). Transmission electron microscope (TEM) imaging shows that the WN/WC–CeO2−x were uniformly confined in the pores of the bowl-shaped carbon matrix (Fig. 2(c)). The corresponding intensity changes in the high-angle annular dark field-scanning transmission electron microscopy (HAADF-STEM) images of WN/WC–CeO2−x@PNC confirmed the formation of a mixture of alloy nanocrystals (Fig. S6, ESI†) and fully exposed single clusters (Fig. 2(d)). Isolated single clusters (marked by red circles in Fig. 2(d)) were clearly visible on the entire support. The HAADF-STEM images of the nanocrystals agglomerated into clusters showed lattice spacings of 0.176, 0.198, and 0.306 nm, corresponding to the distances between the WN(001), WC(100), and CeO2(311) faces, respectively (Fig. S6, ESI†). The HAADF-STEM image of WN/WC–CeO2−x@PNC and the corresponding EDS elemental map (Fig. 2(e)) show that Ce, W, O, N, and P were uniformly distributed on the bowl-shaped carbon. Doping with different heteroatoms can induce, strain, and structural defects in carbon materials,27 and may alter the charge density of the carbon materials,28 thereby potentially improving overall water splitting kinetics.
The normalized X-ray absorption near edge structure (XANES) and the Fourier-transformed extended X-ray absorption fine structure (EXAFS) spectra of W L3-edge and Ce L3-edge were measured to explore the local electron or atomic structure of WN/WC–CeO2−x@PNC. In Fig. 2(f), the W L3-edge threshold E0 of WN/WC–CeO2−x@PNC is situated between that of standard W foil and WO3, indicating an average oxidation state of tungsten ranging between 0 and +6. Fig. 2(g) provides a more detailed insight into the local chemical coordination of tungsten sites within the sample. Here, the reference sample W foil exhibits a prominent peak at ∼2.58 Å (W–W), while commercial WO3 powder shows a primary peak at ∼1.42 Å (W–O). Notably, WN/WC–CeO2−x@PNC displays a main peak at ∼1.63 Å, indicative of the presence of a W–N/C environment. To further validate the aforementioned atomic coordination, Ce L3-edge XAFS and XANES analyses were conducted. As depicted in Fig. 2(i), the scattering paths of the Ce center in WN/WC–CeO2−x@PNC within the R space exhibit a prominent peak associated with Ce–O coordination at ∼1.98 Å. Furthermore, by analyzing the Ce L3-edge XANES spectra presented in Fig. 2(h), the absorption threshold positions for WN/WC–CeO2−x@PNC and pure CeO2 were determined. The slight reduction in the Ce valence of WN/WC–CeO2−x@PNC indicates further electron redistribution characteristics upon the introduction of tungsten.
The powder X-ray diffraction (XRD) pattern of the WN/WC–CeO2−x@PNC sample (Fig. 3(a)) exhibited several characteristic peaks that matched well with several standard cards. The XRD peaks were consistent with WN (JCPDS no. 89-5266), WC (JCPDS no. 51-0939), and hexagonal close-packed CeO2 (JCPDS no. 81-0792). The wide diffraction peak at 23.5° corresponds to the (002) crystal plane of carbon. Note that the heteroatom-doped graphite-like carbon ensures that the WN/WC–CeO2−x hetero-clusters are wrapped on the hierarchical bowl carbon substrate, which simultaneously improves the intrinsic activity, reaction kinetics, and structural stability of the composite.29 The concentration of bulk oxygen vacancies (OVs) and surface structure of the catalyst were studied by Raman spectroscopy. The Raman spectrum of the WN/WC–CeO2−x@PNC (Fig. 3(b)) showed a strong peak near 458 cm−1 and a weak band around 590 cm−1, corresponding to the typical F2g symmetrical vibration (O–Ce–O extension) and defect induced vibration mode (D band, especially oxygen defect-induced band) of CeO2 with a cubic fluorite structure, respectively.30 The wavenumber difference between the oxygen defect and F2g bands is ∼130 cm−1, which indicates that the OVs of the WN/WC–CeO2−x@PNC catalyst is caused by intrinsic defects (Ce4+→ Ce3+).31 However, the transferred electrons are easily trapped in the OVs, and the electron paramagnetic resonance (EPR) spectrum of WN/WC–CeO2−x@PNC shows a strong signal at g = 2.008 (Fig. 3(c)), which also confirms the defect structure of the oxygen vacancy. It is generally believed that the strength ratio between defect-induced vibration mode and F2g mode (i.e. ID/IF2g) is proportional to the concentration of OV in CeO2-based materials.32 Therefore, we further use the relative ratio of defect band to F2g signal to indicate the concentration of oxygen vacancy. The intensity ratio (ID/IF2g = 0.382) of the WN/WC–CeO2−x@PNC catalyst (Fig. 3(b)) was significantly higher than that of WN/WC@PNC (without the characteristic peak of OVs), indicating that the enhanced deformation of the Ce–O–W interface significantly promotes the formation of surrounding defects. Moreover, the oxygen vacancies shorten the electron transfer path and improve the efficiency of charge separation and transfer, thus accelerating the kinetics of the HER.33 As presented in Fig. 3(b), the D-band (1342.5 cm−1) and G-band (1590.0 cm−1) come from the defect/disorder-induced sp3 carbon and the sp2 hybrid carbon with hexagonal bond in the graphite network, respectively.33 The relative intensity ratio (ID/IG) of these two bands is usually used to estimate the degree of graphitization. A smaller ID/IG indicates a higher degree of graphitization of carbon materials.34 For WN/WC–CeO2−x@PNC, ID/IG = 0.957, indicating a high degree of graphitization, suggesting good electrical conductivity. Studies have shown that there are large number of defects in graphitized carbon, which may be the potential active centers of electrocatalysis.35 Additionally, the highly graphitized carbon also makes a great contribution to the long-term durability of electrocatalyst because of its stronger oxidation resistance at high potentials.36
The adsorption–desorption isotherm of WN/WC–CeO2−x@PNC was classified as type IV with an obvious hysteresis in the relative pressure (P/P0) range of 0.5–1.0, indicating mesoporous characteristics and a high Brunauer–Emmett–Teller (BET) specific surface area, which are beneficial for exposing the active sites in the electrocatalytic process (Fig. S7 and Table S2, ESI†). The material was grown using fully exposed WN/WC–CeO2−x hetero-clusters on 3D porous bowl-shaped carbon. Based on our findings, we suggest that the 3D bowl structure is favorable for capturing various reactants and rapidly releasing H2/O2 products from the electrolyte (Fig. S8, ESI†), exposing more active sites, reducing the energy barrier for intermediate formation, and improving the kinetics of the overall water splitting reaction. This is supported by the modified electronic structure (N and P doping), good electronic conductivity, and a bowl structure rich in active sites, which are necessary for realising multifunctional electrocatalysis, rapid electron transfer, and efficient mass transfer.
The catalytically active surface species of WN/WC–CeO2−x@PNC were analysed using X-ray photoelectron spectroscopy (XPS). The C 1s spectrum was deconvoluted into three main peaks at 284.4, 285.0, and 288.3 eV, corresponding to the C–C, C–N, C–O, and CO groups, respectively (Fig. S9a, ESI†). The peaks of W–N, pyridine N, pyrrole N, and graphite N were observed at 398.0, 398.8, 399.6, and 401.3 eV, respectively, in the N 1s spectrum (Fig. S9b, ESI†). The apparent W–N binding energy peak (398.5 eV) in the N 1s spectrum of WN/WC–CeO2−x@PNC, typical of metal nitrides, confirmed the effective modification of the electronic structure of the carbon materials, which facilitates interaction between the heteroatoms and metals.37–39 The water splitting process is further promoted by the combination of pyridine and pyrrole N, and the formation of oxygen intermediates.40 The O 1s spectrum shows signals of metal–oxygen groups (Ce–O), oxygen vacancies, and lattice oxygen at 530.7, 531.3, and 532.0 eV, respectively (Fig. 3(d)). The P 2p spectrum (Fig. S9c, ESI†) was deconvoluted into peaks at 130.0 and 129.2 eV, ascribed to P 2p1/2 and P 2p3/2, respectively. The peak near 35.5 eV in the W 2p XPS profile of WN/WC–CeO2−x@PNC was divided into two peaks (Fig. 3(e)) at 36.5 eV (W–N) and 34.5 eV (W–C). The peak for WN/WC–CeO2−x@PNC was positively shifted by ∼0.25 eV compared to that of WN/WC@PNC, indicating that the coupling between WN/WC and CeO2 changed the chemical composition and charge on the WN/WC surface, that is, the transfer of electrons from the WN/WC surface to the CeO2 surface. The mixed valence state of Ce in WN/WC–CeO2−x@PNC was revealed in the high-resolution Ce 3d XPS profile (Fig. 3(f)). The rapid redox/oxidation cycle of Ce3+ and Ce4+ in CeO2 facilitates the loss of oxygen or electrons from CeO2 to form oxygen vacancies or defects, and the valence state and defect structure of CeO2 are dynamic, making it easier to form favourable active sites for catalysis.41 Among them, the binding energy of Ce 3d5/2 and Ce 3d3/2 at 886.1 and 904.7 eV can be attributed to the Ce3+ of 3d104f1 electronic configuration in CeOx, while the other two peaks at 882.1 and 900.7 eV can be attributed to Ce4+ in CeO2.
It is well known that oxygen vacancies are ubiquitous on the surface of the samples, which can adjust the electronic structure, crystal structure and surface chemistry of the catalyst, thus affecting its conductivity, charge transfer kinetics, and the number of active sites and adsorption/desorption free energy. In total these parameters strongly affect the catalytic activity for HER and OER.42 We speculate that the coordinated unsaturated metal atoms around the OVs can effectively trap and activate the reactants as tentatively shown in Fig. 3(g). Here, a “treasure bowl” type hetero-cluster was designed, where individual active clusters are isolated on the “treasure bowl” (Fig. 3(h)). The synergy of OVs defects would allow movement within their respective “bowls”, while migration across “bowls” is obstructed, achieving the dynamic constraint design of single clusters.
The electrochemical impedance spectra of the WN/WC–CeO2−x@PNC and WN/WC@PNC catalysts were acquired in 1.0 M KOH solution (Fig. 4(d)). The Rct value of WN/WC–CeO2−x@PNC was 22.7 Ω, significantly lower than that of WN/WC@PNC (44.8 Ω) and CeO2 (312.6 Ω). Thus, compared with the other catalysts, electron transfer is fastest at the electrode/electrolyte interface of CeO2–WN/WC@PNC. We attribute the fast electron transfer to the three-phase heterogeneous interface containing oxygen defects, and the amorphous region of the unsaturated metal atoms, which can adjust the local charge density of the optimised coordination environment towards the H* species.44 Secondly, the close vicinity of the core (W–Ce) of the ultra-fine nanoclusters with the N,P-doped carbon shell should further improve the electronic conductivity.43 To better understand the catalytic mechanism, the electrochemical surface area (ECSA) of the various catalysts was estimated based on the electrochemical double layer capacitance (Cdl), which was measured by cyclic voltammetry in the non-faradaic region (Fig. S11, ESI†). The calculated Cdl for WN/WC–CeO2−x@PNC (94.38 mF cm−2) is higher than that of WN/WC@PNC (54.63 mF cm−2), CeO2 (20.39 mF cm−2) and Pt/C (22.50 mF cm−2) (Fig. 4(e)), indicating a larger electrochemical surface area in WN/WC–CeO2−x@PNC, explained by the larger specific surface area, fully exposed clusters, and hierarchical hollow porous structure. In addition to achieving high catalytic activity, another serious challenge for HER electrocatalysts is operational stability, which may prevent large-scale commercialisation. The stability of WN/WC–CeO2−x@PNC during continuous cycling in 1.0 M KOH was studied. In the 20 h chronoamperometric measurement, the current density change is negligible (Fig. 4(f)), and after HER electrocatalysis for 3000 cycles, the LSV curves showed negligible negative shifts compared to those of the fresh catalysts.
Following stability tests, SEM and STEM characterisation (Fig. S12, ESI†) demonstrate a well-preserved morphology and structure of WN/WC–CeO2−x@PNC, in line with the demonstrated good stability. Furthermore, post-durability test XPS analysis and electrolyte ICP-OES testing provide further insights into the catalyst stability and ultra-low leaching, revealing the presence of elements such as W 4f, Ce 3d, N 1s, P 2p, and O 1s, consistent with the expected functionalities (Fig. S13 and Table S3, ESI†). These findings confirm minimal leaching of metal ions and underscore the corrosion resistance of the WN/WC–CeO2−x@PNC electrode in alkaline solutions, particularly in the context of the HER.
The excellent electrocatalytic performance of WN/WC–CeO2−x@PNC can be explained by the variety of favorable structural characteristics at different scales. At the atomic level, the addition of N/P heteroatom can cooperatively change the electronic structure of carbides and improve their intrinsic activity, thereby significantly promoting hydrogen evolution.43 On the nanometer scale, the hierarchical hollow bowl-like structure of WN/WC–CeO2−x@PNC and fully exposed hetero-clusters are beneficial for promoting mass transport and exposure of rich active centers. From the perspective of catalytic kinetics (Fig. 4(g)), that is, molecular diffusion processes (external and internal diffusion of reactants or products) and surface reaction processes (adsorption, surface reaction and desorption), the hollow nanostructures provide potential opportunities for controllable catalysis by purposefully regulating the diffusion, adsorption and surface reactions of the molecules and intermediates. Moreover, with the generation of intrinsically defective OVs, the remaining electrons confined to the OVs flow more easily to the surrounding Ce/W atoms, forming an electron-rich region on the metal atoms.45,46 More precisely, the delocalised electrons around the metal atoms exhibit stronger attraction of hydrogen and promote the adsorption of hydrogen, thus adding more active sites to the surface of the inert base.
To evaluate the bifunctionality of the catalyst for use in next-generation fuel cells and zinc–air/CO2 cells, the OER performance was assessed under the same alkaline conditions. Compared with the overpotential of WN/WC@PNC (382 mV), the overpotential and Tafel slope (99.5 mV dec−1) of WN/WC–CeO2−x@PNC, determined from the LSV curve (Fig. S14a, ESI†), were significantly lower (Fig. S14b and d, ESI†). An overpotential of only 286 mV was required to reach a current density of 10 mA m−2. The introduced CeO2 has redox-active Ce4+/Ce3+ pairs (the electronic structure of Ce is [Xe] 4f15d16s2, where Ce3+ and Ce4+ are stable and interconvert easily). The abundant lattice oxygen vacancies lead to a significant increase in the OER activity via electronic modulation of the active sites.29 WN/WC–CeO2−x@PNC showed excellent stability after 24 h in the chronoamperometric test, with only a slight decrease in the current density (Fig. S14c, ESI†), attributed to good coupling between CeO2 and the tungsten-based transition metal compounds, which produced a unique electrocatalyst with enhanced electronic conductivity and structural stability through the synergistic effect.
WN/WC–CeO2−x@PNC was further used as anode and cathode electrocatalysts to evaluate the overall water splitting performance in a two-electrode cell (Fig. 5(a)). For comparison, an electrolytic cell employing commercial RuO2 and Pt/C as the cathode and anode, respectively, was also tested. The battery voltage with the WN/WC–CeO2−x@PNC electrolytic cell was 1.57 V at 10 mA cm−2, which is better than that obtained with the RuO2 and Pt/C electrode pairs (1.88 V). The evolution of H2 and O2 gas was observed almost instantly upon applying the potential (Fig. 5(a) inset). Notably, the overall hydrolytic performance of this electrolytic cell is better than that achieved with most previously reported non-precious-metal electrocatalysts (Fig. 5(b) and Table S5, ESI†), indicating good potential for practical overall water splitting. In addition, WN/WC–CeO2−x@PNC has excellent durability, after continuous chronopotentiometry measurements at a current density of 50 mA cm−2 (Fig. 5(d)) in 1.0 M KOH for 100 h, the electrocatalytic performance decreased negligibly (the small deviation of the current density may be due to continuous bubble formation on the electrode surface, which blocks the active electrode area). This demonstrates that the WN/WC–CeO2−x@PNC have excellent catalytic performance in alkaline conditions, making them highly promising catalysts for energy storage by hydrogen production from solar energy.
![]() | ||
Fig. 5 Solar-driven water splitting configurations and electrochemical performance under actual photovoltaic conditions. (a) Polarization curves of WN/WC–CeO2−x@PNC for overall water splitting. (b) Cell voltages of different electrolysers constructed with bifunctional electrocatalysts for water electrolysis (Table S3, ESI†). (c) Digital photograph of water electrolyser driven by solar cell integrated system. Inset: Photograph of the as-designed two-electrode system confirming the generation of H2 and O2 gas bubbles during water electrolysis process. (d) Chronoamperometric stability over 100 h at an applied cell voltage of 1.70 V in alkaline medium (1.0 M KOH). (e) Solar-driven water electrolyser system. (f) Current density vs. voltage curve (J–V curves) of commercially available GaAs-solar cell under AM 1.5G illumination and LSV polarization curve of WN/WC–CeO2−x@PNC for overall water splitting (the operating current density is ∼15.38 mA cm−2) (g) Comparison of the solar-to-hydrogen (STH) efficiency achieved in this work with that in the literature (Table S6, ESI†). |
A mechanism for the HER in alkaline electrolytes was proposed, as illustrated in Fig. 6(c). H2O molecules are initially adsorbed on the W sites near the WN/WC–CeO2−x-Vo@PNC interface. Subsequently, the oxygen vacancies (Vo) in CeO2 facilitate the dissociation of H2O into OH* and H*. Hence, this reaction becomes the rate-determining step (RDS), where the generated H* is promptly transferred to and adsorbed on the adjacent Ce–W sites, enabling the subsequent transformation of H to H2. The synergy between WN/WC and CeO2 facilitates the adsorption and desorption of reaction intermediates. Remarkably, WN/WC–CeO2−x@PNC exhibited a lower H* desorption energy barrier (0.21 eV) than WN/WC@PNC (0.49 eV) and WN/WC–CeO2@PNC (0.32 eV), indicating that oxygen vacancies and hetero-clustered WN/WC–CeO2−x disrupt the neutrality of the initial constituents, generating unsaturated atoms and establishing efficient electronic transfer pathways. Ultimately, this orchestrated charge redistribution successfully changes the adsorption energy of the intermediate species in the RDS, thereby catalysing the electrochemical redox reactions.
Footnotes |
† Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d4mh00111g |
‡ These authors contributed equally. |
This journal is © The Royal Society of Chemistry 2024 |