Lijuan
Jiang
,
Ruijing
Wang
,
Huimin
Zhou
,
Guang-Feng
Wei
* and
Xuefeng
Wang
*
Shanghai Key Lab of Chemical Assessment and Sustainability, School of Chemical Science and Engineering, Tongji University, Shanghai 200092, China. E-mail: weigf@tongji.edu.cn; xfwang@tongji.edu.cn
First published on 15th January 2024
Carbides have a high d-band electronic density of states at the Fermi level and hence characteristics resembling those of platinum. However, due to the strong adsorption of hydrogen intermediates on the surface of the carbides, the kinetics of the hydrogen evolution reaction (HER) is severely limited. Here, we demonstrate a strategy to improve the HER kinetics of carbides by modulating the surface electronic structure. Mo2C–W2C ultrafine nanoparticles on reduced graphene oxide (Mo2C–W2C/RGO) with strong electron interaction are constructed to improve the HER activity. The interfacial interaction between Mo2C and W2C adjusts the electronic structure of the surface, and the favorable hydrogen desorption on this surface is proved by density functional theory calculations. Benefitting from the adjusted electronic structures on the Mo2C–W2C heterointerface, the hydrogen adsorption free energies on both the Mo and W sites of Mo2C–W2C are close to the ideal values (0 eV). As expected, the Mo2C–W2C/RGO exhibits a superior HER activity with ultra-low overpotentials of 81 mV and 87 mV at a current density of −10 mA cm−2 under acidic and alkaline conditions, respectively, as well as remarkable stability. This work reveals the modulating effect of the electronic structure of carbides on the HER activity, which has a profound guiding effect for the directional design of advanced electrocatalysts.
Remarkable advances have been made regarding the use of transition metals and their alloys, carbides, nitrides, chalcogenides, and phosphides, which feature their high activity, outstanding stability, and earth abundance.11,12 Among the candidates, transition-metal carbides, such as molybdenum carbides and tungsten carbides, have attracted increasing interest on account of their noble-metal-like electronic configuration, high electronic conductivity, wide pH applicability and outstanding catalytic activity correlated with tunable phase composition.13,14 In addition, a previous report has proposed that the hybridizing between the d-orbital of metallic tungsten/molybdenum atoms and s- or p-orbital of carbon atoms gave rise to a broadened d-band of molybdenum carbides and tungsten carbides resulting in high similarity to the d-band of Pt.15–18 Despite its unique properties, pure carbides show low activity for the HER.19 On the one hand, density functional theory (DFT) calculations have demonstrated that the strong adsorption of H* on the carbide surface facilitates the reduction of H+ but restricts the hydrogen desorption step.20 On the other hand, the synthesis of carbides usually requires a high-temperature annealing process, resulting in serious agglomeration of particles to form an irregular morphology with low specific surface area, thereby reducing the density of the active site.21,22 Efforts have been devoted to designing carbide nanostructures with large surface area and well-exposed active sites. For instance, Hu et al. reported a scalable and economical method to produce W2C nanodot-decorated CNT networks via a spray-drying approach.23 Zheng et al. presented a strategy to synthesize well-dispersed transition metal carbide nanoparticles.24 These strategies have been successfully used to prepare a variety of carbide nanomaterials with effectively improved activity, but the size of the as-obtained nanoparticles is hard to control and the size distribution range is wide. At the same time, these strategies for adjusting the nanoarchitectures of carbides cannot change their inherent properties of the strong adsorption of hydrogen intermediates, which also leads to the limitation of the activity of carbides. To address this issue, numerous approaches are being pursued to modify carbides by constructing heterophases and doping with heteroatoms to adjust the electronic properties and enhance the HER activity.19,20,25,26 Notably, the strong interaction at the two-phase interface can establish electronic transfer and generate electronic density redistribution between the two adjacently contacted components, adjusting the electronic structure and promoting electrocatalytic performance. Chen et al. identified that a synergy between Mo2N and Mo2C in biomass-derived heterostructures can exert the activity and satisfactory durability.25 From this perspective, constructing small-sized carbide nanoarchitectures and modulating the electronic structure of carbides should be an effective way to branch out for studying catalytic abilities.
Herein, a heterostructure composed of Mo2C and W2C anchored to reduced graphene oxide (denoted as Mo2C–W2C/RGO) is constructed via a sequential two-step approach. The Mo2C–W2C heterostructure with ultrafine size (about 2–3 nm) and abundant interfaces maximizes the exposure of active sites and facilitates the mass transfer process, while the interfacial interaction optimizes the electronic configuration of the catalyst. DFT calculations reveal the modulating effect of the electronic structure of the Mo2C–W2C heterostructure on the HER activity. The Mo2C–W2C heterostructure shows a favourable hydrogen desorption step, resulting in the optimized free energy for hydrogen adsorption (ΔGH*). As expected, the Mo2C–W2C/RGO affords an ultra-low overpotential (η10 = 81 and 87 mV for reaching a current density of −10 mA cm−2) under acidic and alkaline conditions.
The Gibbs free energy changes (ΔG) for hydrogen adsorption were expressed as:
The XRD investigation confirmed the formation of Mo2C–W2C/RGO heterostructures in the final products (Fig. 3a), and the results of Mo2C/RGO, W2C/RGO and Mo2C–W2C/RGO show the characteristic pattern of hexagonal W2C (JCPDS no. 35-0776) and Mo2C (JCPDS no. 35-0787). The diffraction peaks of all samples are relatively weak and broad, consistent with the ultrafine nanoparticles shown by SEM and TEM results. Accordingly, these samples present two distinct characteristic peaks of the carbon-based networks, located in the D (1350 cm−1) and G (1580 cm−1) bands in Raman spectra (Fig. 3b).35,36 The IG/ID value for GO sheets (1.25) is significantly higher than that for RGO nanosheets (0.98) because of the decrease in oxygenic groups during the thermal treatment (Fig. S3†). In contrast, the IG/ID values of Mo2C–W2C/RGO, W2C/RGO, and Mo2C/RGO are markedly increased compared with RGO, which could be due to the coupling between carbide nanoparticles and RGO resulting in a more disordered carbon structure.
![]() | ||
Fig. 3 (a) XRD patterns and (b) Raman spectra of the Mo2C/RGO, W2C/RGO and all Mo2C–W2C/RGO samples. XPS profiles of (c) Mo 3d and (d) W 4f in the Mo2C/RGO, W2C/RGO and all Mo2C–W2C/RGO samples. |
To analyze the surface chemical composition and electronic structure of the samples, X-ray photoelectron spectroscopy (XPS) was performed. As depicted in Fig. S4a,† an XPS survey was used to identify the composition of the Mo, W, and C elements in these samples. The profiles of Mo 3d in Mo2C–W2C/RGO and Mo2C/RGO can be deconvoluted into six peaks (Fig. 3c), suggesting that there are three species (Mo–C, Mo4+, and Mo6+) for Mo on the surface.37,38 For the Mo2C/RGO sample, the couple at 228.5 and 231.5 eV should be ascribed to the Mo 3d5/2 and 3d3/2 of Mo–C species. The peaks located at 229.2 and 232.9 eV are attributed to Mo4+ species, and the two at 232.2 and 235.5 eV are associated with Mo6+ species. They are commonly observed when carbides are exposed to air.39 The W 4f spectrum of W2C/RGO (Fig. 3d) can be fitted into four peaks with binding energies at 35.5 and 37.7 eV corresponding to W 4f7/2 and 4f5/2 of surface oxidized W6+, while the two peaks at 31.8 and 33.8 eV can be attributed to W 4f7/2 and 4f5/2 of W–C species.40 Similarly, Mo 3d and W 4f for all Mo2C–W2C/RGO samples can also be fitted into corresponding peaks. Regarding the consistent binding energy of the oxidation state affected by air (Mo6+, Mo4+ and W6+) in the peak fitting of all the samples, we focus on the binding energy of Mo–C and W–C species affected by the heterostructure, which are regarded as the active centers for electrocatalytic HER.40,41 The Mo2C/RGO sample presents Mo 3d5/2 and 3d3/2 at 228.5 and 231.5 eV for Mo–C species, respectively. By contrast, these peaks for Mo–C species slightly shift toward lower binding energies in all Mo2C–W2C/RGO samples, indicating that a change in the electronic structure occurs due to introduction of the W2C. On the other hand, the binding energy of the W–C species of Mo2C–W2C/RGO is positively shifted from that of W2C/RGO samples. Given the higher electronegativity of Mo (2.16) compared to W (1.7),42 the negative shift of Mo–C and positive shift of W–C species in all Mo2C–W2C/RGO samples can reasonably be attributed to electron transfer from W to Mo. Notably, the binding energies of Mo shifted negatively with increasing W composition in Mo2C–W2C/RGO, suggesting that the electronic structure of the surface of the samples can be effectively modified by varying the Mo/W ratio (Fig. 3c and d). The above high-resolution XPS analysis shows that the construction of the Mo2C–W2C/RGO heterostructure modulates the electronic structure of the original carbide, which is expected to regulate the adsorption/desorption strength between active sites and adsorbates. Fig. S4b† shows the high-resolution C 1s XPS spectrum for the all samples, which can be deconvoluted into five peaks (C–C/CC, C–N, C
O, O
C–O, and C–Mo/C–W).43,44 Moreover, the N 1s XPS profile (Fig. S4c†) can be fitted into four peaks, assigned to Mo 3p, pyridinic-N, pyrrolic-N, and quaternary-N, respectively.45–47 In general, the planar structures of pyrrolic-N and pyridinic-N favor electron transport in π–π-conjugated graphite layers.40 Regarding the similarity in N type and content, the contribution to conductivity by N-doping into the carbon matrix can be considered to be equivalent for W2C/RGO, Mo2C/RGO, and all Mo2C–W2C/RGO samples.
The electrocatalytic HER performances of the samples were first evaluated in the standard three electrode system under 0.5 M H2SO4 conditions at 5 mV s−1. All the samples are supported on glassy carbon electrodes with the same mass loading of 0.24 mg cm−2. The synthesis conditions of different molar ratios of W and Mo are at first screened (Fig. S5†), and all the Mo2C–W2C/RGO materials are superior to the W2C/RGO and Mo2C/RGO (Fig. S5 and Table S2†). At the same time, it is suggested that the optimal electrocatalyst with an appropriate Mo/W molar ratio is Mo2C–W2C/RGO-0.56. Fig. 3a displays the polarization curves of the W2C/RGO, Mo2C–W2C/RGO-0.56 and Mo2C/RGO, along with bare RGO and the commercial Pt/C (20 wt%) benchmark for reference. It is obvious that RGO has a negligible HER activity. The W2C/RGO and Mo2C/RGO afford overpotentials (η10) of 158 mV and 110 mV to achieve a current density (j) of −10 mA cm−2, and Mo2C–W2C/RGO-0.56 exhibits an ultralow η10 of only 81 mV. This result reveals that the unique two-phase heterostructure under an optimal Mo/W ratio can dramatically boost the catalytic activity. In addition, the catalytic performance of Mo2C–W2C/RGO-0.56 is much better than those of molybdenum- and/or tungsten-based multi-phase heterostructures at −10 mA cm−2, such as twinned WCN (128 mV),48 W2C/WP@NC (116.37 mV),49 Mo2N–Mo2C/HGr (157 mV),50 and WC-W2C/HCDs (96 mV),51 and is the best among the non-noble catalysts reported recently (Table S3†).
To further investigate the HER mechanism, the Tafel plots (Fig. 4b) were derived by plotting the overpotential against log(j) according to the Tafel equation (η = b × log(j) + a).52 Mo2C–W2C/RGO-0.56 delivers a low onset overpotential (ηonset) of 30 mV and a low Tafel slope (b) of 56 mV dec−1, superior to those of Mo2C/RGO (ηonset = 43 mV, b = 73 mV dec−1) and W2C/RGO (ηonset = 67 mV, b = 94 mV dec−1). Generally, the HER process includes two steps, namely the Volmer, Heyrovsky or Tafel steps,1,53 corresponding to Tafel slopes of 120, 40, and 30 mV dec−1, respectively. The above results suggest that the HER kinetics for Mo2C–W2C/RGO-0.56, Mo2C/RGO and W2C/RGO should follow the Volmer–Heyrovsky mechanism and Mo2C–W2C/RGO-0.56 suggests a more rapid HER kinetics.54,55 Extrapolating the above Tafel plots, the higher value of the calculated exchange current density (j0) also indicates the high HER activity of Mo2C–W2C/RGO-0.56 (Table S4†).
The electrochemically active surface area (ECSA) of the samples was estimated by the double layer capacitance (Cdl).56,57 In accordance with the obtained Cdl values (Fig. S6†), Mo2C–W2C/RGO-0.56 (38.5 mF cm−2) has the highest ECSA of all samples, thereby further confirming the higher performance of Mo2C–W2C/RGO-0.56. In addition, the electrochemical impedance spectroscopy (EIS) measurements (Fig. 4c) show that Mo2C–W2C/RGO-0.56 has the smallest electron transfer resistance and the EIS trend is in good agreement with the HER activity trend. The HER activity of the above samples along with that of the commercial Pt/C catalyst is summarized in Table S4.†
As another important assessment of the electrocatalyst, a long-term electrochemical stability test of Mo2C–W2C/RGO-0.56 in acidic media was also performed (Fig. 4d). The activity of the working electrode remained almost completely stable in the continuous test of 30 h, and the polarization curve of the catalyst after the stability test showed no measurable decline of the cathodic current (inset of Fig. 4d). Accordingly, SEM investigation confirmed the well-retained nanostructures after the stability test, as shown in Fig. S7.† These results demonstrate that Mo2C–W2C/RGO-0.56 possesses outstanding durability in acidic electrolyte.
Additionally, Mo2C–W2C/RGO-0.56 also exhibited good HER performance in the alkaline medium (Fig. 5 and Table S5†). Apparently, the performance improvement after the formation of the Mo2C–W2C heterostructure can be obviously observed from Fig. S8 and Table S6.† By contrast, Mo2C–W2C/RGO-0.56 displays a relatively small overpotential of 87 mV at −10 mA cm−2, lower than those of Mo2C/RGO (134 mV), W2C/RGO (159 mV), and RGO (491 mV). Besides, commercial Pt/C still presents the best HER performance, requiring an overpotential of merely 27 mV to deliver a current density of −10 mA cm−2. The faster HER kinetics of Mo2C–W2C/RGO-0.56 can be reflected by its small Tafel slope of 59 mV dec−1 and low onset overpotential of 33 mV, lower than the values of Mo2C/RGO (ηonset = 66 mV, b = 64 mV dec−1) and W2C/RGO (ηonset = 76 mV, b = 93 mV dec−1) (Fig. 5b and Table S5†). Besides, the Tafel slope of Mo2C–W2C/RGO-0.56 within the theoretical value indicates that the reaction kinetics for hydrogen generation is the Volmer–Heyrovsky mechanism.54,55 Furthermore, capacitance values (Fig. S9 and Table S5†), Nyquist diagram (Fig. 5d), and j0 (Table S5†) confirm that the Mo2C–W2C/RGO-0.56 electrode has the largest electrochemical surface area, the smallest charge transfer impedance, and higher intrinsic activity, which is consistent with its HER activity. At the same time, Mo2C–W2C/RGO-0.56 also has satisfactory long-term stability in 1.0 M KOH, delivering a current density of about −10 mA cm−2 for 30 h. In addition, SEM analysis results (Fig. S10†) prove that the structure of Mo2C–W2C/RGO is well maintained after a long-term HER process, which further confirms the remarkable stability of Mo2C–W2C/RGO. Remarkably, the Mo2C–W2C/RGO delivers prominent HER activity in 1.0 M KOH, representing one of the current best carbide-based HER catalysts (Table S3†). All of these results indicate that Mo2C–W2C/RGO shows excellent hydrogen production performance over the entire pH range, providing a promising HER electrocatalyst for practical application.
The aforementioned results indicate the formation of the Mo2C–W2C heterostructure coupled with RGO, which shows superior acidic and alkaline HER performance. The origin of the efficient HER activity of Mo2C–W2C/RGO is worthy of investigation. Therefore, a series of control samples, including W2C, Mo2C, and Mo2C–W2C (Fig. S11a, c and e†), were constructed in the absence of GO, in addition to the RGO, W2C/RGO and Mo2C/RGO already mentioned. Electrochemical tests were performed to compare the activity of the HER under acidic and alkaline conditions. The poor performance of the RGO indicates that RGO is not an active component (Fig. S12†). In addition, W2C/RGO and Mo2C/RGO show superior HER activity over W2C and Mo2C, respectively, and the HER activity of Mo2C–W2C/RGO is better than that of Mo2C–W2C. The results indicate the positive role of RGO in the improvement of the HER performance. This is related to the good electrical conductivity and large surface area of RGO. In the case of the catalysts without a graphene substrate, W2C, Mo2C, and Mo2C–W2C catalysts exhibit severe agglomeration and accumulation (Fig. S11b, d and f†). The good coupling between graphene and nanoparticles inhibits the agglomeration of nanosheets and plays a good dispersion role, which helps to improve the active surface area of the catalyst and expose the active site. Furthermore, graphene as a good conductor can effectively reduce the interfacial barrier of electron transfer, which is helpful for the enhancement of reaction kinetics. The improved activity of Mo2C–W2C/RGO over Mo2C/RGO and W2C/RGO implies the important role of the heterostructure for the HER due to its strong electronic interaction. Thus, we propose that the Mo2C–W2C heterostructure is the main active component, and RGO is also an important factor because of its role in exposing the active site, as well as in mass/charge transfer.
To gain a fundamental understanding of the high catalytic activity of the Mo2C–W2C heterostructure for the HER, density functional theory (DFT) calculations were performed. Here, the free energy profiles of H adsorption on the (001) surface of pure W2C (Fig. 6a) and Mo2C (Fig. 6b), along with the corresponding surfaces of Mo2C–W2C (Fig. 6c) were computed. Regarding the overall HER rate, the hydrogen adsorption free energy (ΔGH*) is regarded as a good descriptor to evaluate the activity of the material, and optimal HER catalysts should have ΔGH* close to zero.58–62 As illustrated in Fig. 6d, on the W site of W2C, the ΔGH* is calculated to be −0.32 eV, and the ΔGH* on the Mo site of Mo2C is −0.10 eV. The relatively negative ΔGH* on the W site shows that the strong adsorption of H on the surface of W2C restricts the Heyrovsky process, resulting in a low HER efficiency, which is consistent with our experimental results of electrochemical tests. For Mo2C–W2C, the ΔGH* values on the Mo and W sites of Mo2C–W2C are −0.03 eV and −0.25 eV, respectively, both of which are closer to the ideal ΔGH* (0 eV) than the corresponding Mo and W sites on pure Mo2C and pure W2C. The calculations demonstrate that the interface structure of Mo2C–W2C plays a crucial role in its excellent HER activity.
Additionally, we have calculated the charge-density differences (CDD) as well as the planar-average charge-density difference (PA-CDD) along the Z direction of the heterostructure to evaluate interfacial charge-transfer processes. As shown in Fig. 6e, the electron accumulation (yellow) and depletion (cyan) are concentrated at the interfaces formed between Mo2C and W2C, indicating the strong charge redistribution at the interface, consistent with the electron modulation observed by XPS (Fig. 3c and d). Fig. S13† displays the total density of states (DOS) of the Mo2C, W2C and Mo2C–W2C. The DOSs of all three materials cross their Fermi level (corresponding to the energy of the highest electron occupied orbital of the material), demonstrating that all of them can possess good conductivity during the HER. The high density of states of Mo2C–W2C adjacent to the negative side of the Fermi level further proves its excellent ability to donate electrons during the electrocatalytic HER process. The better HER activity of Mo2C–W2C may also originate from the interface charge distribution. This charge redistribution adjusts the surface electronic structure of the material, which weakens the H adsorption and facilitates H2 desorption resulting in enhanced HER kinetics and reduced overpotential.
Based on all the above experiments and theoretical calculation results, the remarkable electrocatalytic activity and stability of Mo2C–W2C for the HER can be attributed to the following reasons: (i) the material is supported by graphene sheets with good electrical conductivity and large specific surface area, showing ultra-fine and evenly distributed nanoparticle morphology, high ECSA and good electron transport capacity, which is conducive to the transport of reactants and products, thus accelerating the occurrence of the HER. (ii) The optimized electronic structure at the Mo2C–W2C interfaces improves the kinetics of the hydrogen evolution process, which endows the Mo2C–W2C with enhanced kinetics and high activity.
Footnote |
† Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d3ta07373d |
This journal is © The Royal Society of Chemistry 2024 |