Ya-Nan
Zhou
a,
Meng-Xuan
Li
a,
Zhuo-Ning
Shi
a,
Jian-Cheng
Zhou
a,
Bin
Dong
*a,
Wenchun
Jiang
b,
Bin
Liu
a,
Jian-Feng
Yu
a and
Yong-Ming
Chai
*a
aState Key Laboratory of Heavy Oil Processing, College of Chemistry and Chemical Engineering, China University of Petroleum (East China), Qingdao 266580, PR China. E-mail: dongbin@upc.edu.cn; ymchai@upc.edu.cn; Fax: +86-532-86981156; Tel: +86-532-86981156
bCollege of New Energy, China University of Petroleum (East China), Qingdao, 266580, PR China
First published on 25th February 2022
Developing high-density and uniform crystal–amorphous interfaces is highly desirable for the hydrogen evolution reaction (HER). Herein, crystal–amorphous NiO/MoO2 with a coupled high-density interface has been designed to tailor the charge distribution to lower the reaction energy barrier for the HER. The obtained self-supported NiO/MoO2-100-2, which is fabricated through a simple and versatile anodizing-assisted molten salt with a MoNi substrate, possesses high-density and a well-dispersed NiO crystal /amorphous MoO2 heterojunction, benefiting local electron rearrangement and charge transfer. The NiO crystals are beneficial for the alkaline water dissociation to rapidly generate active hydrogen (Volmer step), therefore facilitating the subsequent Heyrovsky and Tafel steps to occur in the amorphous MoO2 region. NiO/MoO2-100-2 exhibits superior HER activity with a low overpotential of 48 mV at 10 mA cm−2, a small Tafel slope of 51.5 mV dec−1 and robust stability, which can be chalked up to the more available actives sites, enhanced conductivity and favorable H adsorption sites derived from the modulated charge and d-band structure near the Fermi level.
Transition metal oxides are generally regarded as HER inactive materials due to unsuitable hydrogen adsorption energy and inert electron transport.7–9 Particularly for nickel oxide (NiO), which is an ideal catalyst for water adsorption and dissociation in an alkaline medium, the calculated ΔGH* value is −0.27 eV or −0.55 eV,10,11 which is too negative to easily desorb active hydrogen, while according to the Sabatier principle, better catalytic performance is based on moderate bonding to the active species.12,13 Therefore, combining NiO with materials with better hydrogen desorption capability to generate efficient catalytic sites for fast hydrogen adsorption/desorption is necessary. Encouragingly, molybdenum dioxide (MoO2) featuring a metallic character possesses a ΔGH* value of around 0.15 eV, exhibiting outstanding potential as a hydrogen adsorption promoter, thanks to the lower unoccupied orbital in the alkaline electrolytes.14–16 Herein, the coupling of NiO and MoO2 is a promising strategy for better HER efficiency. However, the inherently poor conductivity and activity of NiO and MoO2 remain the primary challenge. Intrinsically, its motivation lies in local electron configuration arrangement to optimize the electron transport capacity together with adsorption/dissociation of H2O and hydrogen, which are intertwined with the formed lattice, charges, and spin ordering in obtained catalysts. For the HER, the electron distribution and band structure around the Fermi energy level are extremely remarkable characteristics.17–19 In this regard, the tailoring for spin ordering and orbital filling of metal ions is perceived as a brand new and efficient approach. Taking all these into account, generating a new strong coupling crystal interface with a unique atomic and electronic structure is expected to bridge the gap between NiO, MoO2 and favorable catalytic activity.
Preferentially, the development of a unique crystal–amorphous NiO/MoO2 interface with high density and good distribution is more attractive while challenging because the extraneous Ni and Mo source cannot be easily mixed uniformly. To solve this problem, the MoNi foam can be applied as the substrate to provide a well-mixed Mo and Ni source, thus the density and distribution of the interface may be ensured, whereas high-strength substrates characterized by the dense internal atomic arrangement or/and a glazed surface, including Mo, W, Ta, and their respective alloys, cannot be treated easily by conventional methods, such as solvothermal reaction and electro-deposition.20 Prospectively, molten salt synthesis has emerged as a facile and environmentally friendly method. For example, Li et al. reported a molten salt NaNO3 system, with which ultra-thin RuO2 nanosheets are prepared successfully for water oxidation.21 Wang's team also employed the molten salt method to disperse Ni ions on TiO2.22 Unlike ordinary solvents, the high solubility, space confinement effect, and strong polarization provided by this special ionic liquid will inevitably have a profound impact on the morphology and electrochemical properties of compounds.23 Given these, the molten salt is necessitated to activate MoNi substrates to obtain highly active and stable catalysts; in the present work, it is referred to as NiO/MoO2.
Parallel to the desirable active interface, the generation of abundant defects, even amorphous phases by coupling NiO and MoO2 is equally intriguing because of a larger proportion of unsaturated metal sites and randomly oriented bonds. Concretely, the structure adaptability of amorphous phases in the short-range can expedite the charge transfer between active sites and intermediates as well as contribute to spin-state manipulation, which is supposed to synergistically increase hydrogen evolution with the NiO/MoO2 interface.24,25 The mutual coupling of strategies enables the simultaneous optimization of multiple constraints to boost hydrogen evolution.
Inspired by all the above discussion, a self-supported crystal–amorphous NiO/MoO2 heterointerface directly derived from MoNi foam with a high-density interface has been prepared via a versatile molten-salt method assisted by anodizing. Scheme 1 schematically shows the synthesis process. The pre-anodizing of MoNi foam produces partly oxidized and stripped NiO/MoO2 species and disturbed surface structure (NiO/MoO2-100), which is conducive to subsequent molten salt treatment and morphological transformation. After being etched by molten salt, NiO/MoO2-100 becomes metastable and meanwhile forms a high-density NiO crystal/amorphous MoO2 interface on the brand new stripped layered structure, which is proved to facilitate electron transfer and intermediate adsorption/desorption with a low overpotential of 48 mV at 10 mA cm−2 and a small Tafel slope of 51.5 mV dec−1. DFT calculation also confirms the regulated charge arrangement near the Fermi energy, which is accompanied by the magnetic moment change. More importantly, the long-term durability further verifies the practicability of the molten-salt method, which provides new possibilities for high-performance catalyst design.
For comparison, Ni/Mo-h was synthesized by the same method as NiO/MoO2-s-h using MoNi foam without anodizing.
Fig. 1 Structure characterization of NiO/MoO2-100 and NiO/MoO2-100-2. SEM images of (a) NiO/MoO2-100 and (b and c) NiO/MoO2-100-2. (d) XRD pattern and (e) particle distribution of NiO/MoO2-100-2. |
These activated surfaces with perturbed atomic arrangements are conducive to subsequent molten-salt method treatment. As expected, with two hours of immersion in molten NaCl (NiO/MoO2-100-2), the closely packed irregular diamond-like nanoparticles were easily detected, most of which are 600 nm in diameter (Fig. 1 and Fig. S2†). The significantly smaller nanoparticles undoubtedly contribute to the increase in the specific surface area, therefore creating enough active sites to promote hydrogen production. The crucial role of anodizing is further corroborated by Ni/Mo-2 samples (Fig. S3†), whose average larger-diameter (1200 nm) nanoparticles are merely the products of further growth of original clusters on pure MoNi foam with slight integration, unlike NiO/MoO2-100-2, which is the product of the relatively thorough post-integration growth. X-ray diffraction (XRD) is performed to identify the samples’ surface composition. Fig. S4a† first confirms that the main phase of MoNi foam is metallic Ni (JCPDS: 70-0989), while peaks of Mo species cannot be clearly discerned due to the shielding effect of the stronger peaks of Ni. For the resulting NiO/MoO2-100-2 (Fig. 1d), strong peaks at 37.1°, 43.2°, 62.9°, 75.4° and 79.4° match well with the (111), (200), (220), (311) and (222) crystal planes of NiO (JCPDS: 47-1049); and the signals of the (100), (−102), (122) and (−411) planes of MoO2 (JCPDS: 78-1071) can also be roughly detected at 18.4°, 31.6°, 63.8° and 73.0°. In addition, the MoO3 species can also be indexed at 27.4°, 29.6°, 35.6° and 42.4° (JCPDS: 01-0706) due to surface oxidation. These results suggest the coexistence of NiO and MoOx components. In addition, the broad peak may result from finely dispersed nano-sized particles or low crystallinity. Compared with NiO/MoO2-100 (Fig. S4b†), whose XRD peaks also show the presence of NiO, the diffraction pattern of NiO/MoO2-100-2 becomes diffuse and wider, which demonstrates more defects after molten salt treatment.
Interestingly, the TEM image of NiO/MoO2-100-2 stripped from the substrate shown in Fig. 2a and b clearly manifests the two-dimensional (2D) morphology, implying a previously well-verified phenomenon that during the reaction, molten NaCl can be inserted into NiO/MoO2 to form a 2D layered material.23,26 High-resolution TEM (Fig. 2c and d) confirms the presence of abundant NiO crystal/amorphous MoO2 heterointerfaces thanks to well mixed Mo and Ni in MoNi foam, where two lattice fringe spacings can be observed, belonging to the d-spacing of 0.209 nm for the (200) crystal plane of NiO and 0.284 nm of the (−102) plane for monoclinic MoO2, respectively, consistent with XRD analysis. In detail, Fig. S5† clearly shows the NiO/MoO2 interface, where the detected 0.209 and 0.241 nm belong to NiO while the ambiguous crystals with the distance of 0.218 and 0.147 nm come from MoO2.
The formation of the distinct crystals/amorphous in this work is mainly because of the unique MoNi foam, which is manufactured by deposition with a polymeric sponge as the substrate to form the Ni film–Ni coating–NiMo alloy structure. The Mo and Ni atoms in sandwich-like structures suffer from different moving rates to lead to the dense distribution of the heterophase. According to the earlier report, NiO can facilitate the dissociation of water in an alkaline electrolyte to generate active hydrogen, while amorphous areas are the active centers for hydrogen adsorption and desorption, thus synergistically boosting the HER process.10,11,27 It is worth mentioning that the lattice plane of MoO3 cannot be well observed due to the multi-defect structure. Note that the low valence-state Mo species are deemed to be active for the HER, and can serve as catalytic sites for hydrogen adsorption.14–16 Furthermore, SEM mapping (Fig. 2e) shows that Ni, Mo, and O are uniformly distributed in the selected field. Energy-dispersive X-ray spectroscopy (EDX) results in Fig. S6† show the elementary composition and contents. Notably, it shows a much lower Mo (2.64 At.%) content than Ni content (68.43 At.%), which may result from the different dynamic migration rates of Ni and Mo from the interior of the MoNi foam substrate to the surface in the molten salt system, which can be laterally confirmed by the higher compactness of the pure molybdenum plate than that of nickel foam. Next, the possible growth mechanism of the nanosheet structure is discussed and proposed as follows. First, according to the Arrhenius equation,28 the mathematical relationship of ka and kb that represents the reaction rate constant of supposed growth direction [001] and [100], respectively, is first expressed as below:
Further, X-ray photoelectron spectroscopy (XPS) is employed to investigate the surface composition of NiO/MoO2-100-2. The full scan spectrum confirms the presence of Ni, Mo, and O elements again (Fig. S7a†). The deconvoluted Ni 2p region contains two pairs of typical peaks accompanied by the satellite peaks (labeled as Sat.) (Fig. 3a), which can be reliably assigned to the NiO (2p3/2 at 856.1 eV and 2p1/2 at 873.4 eV) and Ni0 (2p3/2 at 854.1 eV and 2p1/2 at 871.6 eV) from the substrate.33–36 In the Mo 3d spectrum (Fig. 3b), the coexistence of Mo3+, Mo4+ and Mo6+ is substantiated by peaks at 231.9 eV, 232.3 eV and 235.6 eV, respectively.37,38 Moreover, the perturbed electronic structure on the surface of NiO/MoO2-100-2 via the molten-salt method can be discerned by the slight negative shift of Ni 2p (Fig. S7b†) and positive shift of Mo 3d peaks compared with the NiO/MoO2-100 precursor. The O 1s spectrum can be fitted into lattice oxygen (O1 at 529.7 eV), oxygen defects (O2 at 530.2 eV), and surface-adsorbed oxygen species (O3 at 531.7 eV), respectively (Fig. S7c†).39,40 The concentration of oxygen defects in the final catalyst increases to 35.6% from 31.4% of the precursor. The accumulated oxygen vacancies are conducive to electron rearrangement by acting as electron capture sites to regulate the coordination environments and electronic states of surface adsorbents. Meanwhile, the oxygen vacancies can activate their adjacent oxygen atoms toward hydrogen adsorption and facilitate water dissociation, in turn substantially improving the HER performance.41 These XPS results suggest the local rearranged electronic structure of NiO/MoO2-100-2 after the molten-salt method treatment.
The HER performance of all the prepared samples was also measured through three-electrode configuration in 1 M KOH. Fig. 4a shows the related polarization curves, and the target NiO/MoO2-100-2 exhibits the best HER activity with a required overpotential of 48 mV at 10 mA cm−2, superior to NiO/MoO2-100 (204 mV), Ni/Mo-2 (198 mV) and other reported NiO-based catalysts. Intriguingly, when the current density is above 18 mA cm−2, its activity is even better than that of the commercial Pt/C benchmark catalyst. The superiority of the NiO/MoO2 heterostructure is also highlighted by the lower HER activity of single NiO and MoO2 in Fig. S8.† Expectedly, the ratio of MoO2 to NiO has a significant impact on the HER activity, which is determined by the anodizing and molten-salt process. Thus, the influence of these two variables on the ratio of MoO2 and NiO and catalytic performance has been further investigated. In comparison, various anodizing times (50 s, 100 s, 150 s, and 200 s) are researched to clarify the role of pre-oxidation. As shown in Fig. S9a,† the HER performance of control samples increases first and then decreases rapidly with increasing anodic oxidation time, and the optimal time is 100 s, which can be reasonably attributed to the inadequate/excessive activation that leaves the induced NiO/MoO2 inappropriate (too much MoO2) for hydrogen evolution. Similarly, this reason can be applied to the phenomenon after changing the molten-salt method time, which results in excessive production of NiO (Fig. S9b†). Based on these, we speculate that the induced electron redistribution of the NiO/MoO2 heterointerface in the molten salt system is also accountable for the different HER performance. Besides, the water oxidation activity of the best-performing NiO/MoO2-100-2 is also acceptable (Fig. S10†). The corresponding Tafel slopes calculated from polarization curves are depicted in Fig. 4b to grant access to the HER kinetic process.42 The much smaller Tafel slope of 51.5 mV dec−1 for NiO/MoO2-100-2 than NiO/MoO2-100 (198.7 mV dec−1), Ni/Mo-2 (180.2 mV dec−1), Pt/C (56.7 mV dec−1) and other control materials indicates that the rate-determining step changes from the Volmer step to Volmer–Heyrovsky step.43 The influence of time of the molten-salt method and anodizing time exerted on the electron structure can be perceived by a series of Tafel plots (Fig. S11†). Furthermore, double-layer capacitance (Cdl) is evaluated by a range of cyclic voltammetry (CV) curves at 0.21 V vs. RHE (Fig. S12†) to assess the electrochemical active surface area. As expected, NiO/MoO2-100-2 has the highest Cdl value (Fig. 4d), indicating that the increased active sites play a significant role in activity enhancement. In order to set forth the contribution of the surface area and intrinsic activity of samples to the HER, we also normalized polarization curves by the electrochemical active surface area, as shown in Fig. S13.†NiO/MoO2-100-2 still shows admirable HER activity compared to the control sample NiO/MoO2-100 and Ni/Mo-2, indicating that the inherent activity of NiO/MoO2-100-2 improves due to the more intimate interactions between Ni and Mo caused by molten-salt treatment.
Fig. 4 Electrochemical measurements of the obtained NiO/MoO2-100-2 sample in 1 M KOH. (a) Polarization curves and (b) the corresponding Tafel plots. (c) Comparison of the overpotential at 10 mA cm−2 and Tafel slope between NiO/MoO2-100-2 and the reported NiO-based catalysts (Table S1†). (d) Cdl value. (e) Nyquist plots fitted to the model inset. (f) Polarization curves of NiO/MoO2-100-2 before and after 4000 CV scans. The inset is the chronopotential curve at 100 mA cm−2. |
As a prevailing criterion for charge-transfer kinetics of the electrocatalytic process, electrochemical impedance spectroscopy (EIS) data are fitted via the inset circuit model shown in Fig. 4e and Fig. S14.† Obviously, as reflected from the semicircle diameter, the charge-transfer resistances of NiO/MoO2-100-2 are much lower than that of other samples, signifying the favorable charge transfer capability. In order to obtain an insight into the stability of NiO/MoO2-100-2 in the alkaline HER, the polarization curves before and after 4000 CV cycles are first collected in Fig. 4f and no distinct difference can be observed. Moreover, electrocatalytic HER is allowed to proceed for 40 h at a current density of 50 mA cm−2 with negligible degradation (illustrated in Fig. 4f). Fig. S15† shows that the morphology after the stability test remains unchanged and the crystal plane of NiO is well preserved, which together suggests the admirable durability benefitting from the inherent strong adhesion of the NiO/MoO2-100-2 catalyst using the substrate as the metal source. XRD also discerned the NiO and MoO2 phases after the 40 h HER tests (Fig. S16†), while the wider XRD peaks may indicate that the catalyst surface also consists of a NiO/MoO2 heterostructure with more defects. The Ni 2p XPS shifts to a lower bonding energy, meanwhile the Mo 3d XPS moves to higher bonding energy, and no new valence can be deconvoluted (Fig. S17a–c†). For the O 1s spectrum, the content of defective O further increases to 43.2%, implying the newly produced oxygen defects, which is in line with the XRD result. These characterization studies after HER measurement indicate that the surface of NiO/MoO2-100-2 is partially consumed under such a negative potential and alkaline electrolyte, and the newly formed defective species can still serve as highly efficient active centers.
Furthermore, to gain a deeper insight into the significance of the integration of NiO and MoO2, it is necessary to figure out the spin state. According to the crystal field theory, when the electrons rearrange, the splitting of energy levels would result in additional energy. The Mo atom owns the 4d5 orbital with a higher energy level and it is a priority to be in a low spin state. Based on this, the possible electronic structure of MoO2 can be deduced as shown in Fig. 5a. The ligand field of Mo4+ can split the 4d orbital into t2g and eg sets, therefore the two electrons are prone to lie in the t2g orbit. Due to the asymmetric occupation in degenerate orbitals, the two electrons occupy the t∥ orbital, resulting in a lower energy level.44 Similar splits can happen to Ni2+, while the orbit structures of t62ge2g, and eg orbitals have been equally occupied, thus there is no further splitting (Fig. 5b).
In a nutshell, MoO2 and NiO both possess lone pair electrons. Previous reports have indicated that the lone pair electrons of nanomaterials can induce strong interactions between them, thus improving reactivity and stability via partial or entire electron-donating.45 In the present work, by means of a unique molten method, NiO can closely couple with MoO2 on the interface of NiO/MoO2, and the fluidity of unpaired electrons may lead to stronger electrical perturbations to enhance HER performance. Furthermore, the density of states (DOS) and partial DOS (PDOS) of the prepared catalysts were investigated within the density functional theory to verify the vital role of the NiO/MoO2 heterostructure for hydrogen evolution. The possible structures of NiO, MoO2 and NiO/MoO2 have been constructed and optimized, as shown in Fig. S18.† The calculated DOS shown in Fig. 5c suggests the favorable conductivity of NiO/MoO2 materials, and the d-electrons of Ni and Mo make the main contributions to the electron arrangement near the Fermi level. Besides, the level of the d-band center was also studied to quantify intuitively the electronic modulation.46 As Fig. 5d shows, the d-band center energies of NiO and MoO2 are −0.90 eV and −3.39 eV, respectively, which is too close or too far from the Fermi level, indicating that the metal–H bond is too strong or too weak to effectively desorb/trap hydrogen protons, thus resulting in low HER activity. Meanwhile, for the heterogeneous interface of NiO/MoO2, the d-electrons of the two rearrange and the d-band center energy is −1.02 eV, falling somewhere in between and manifesting moderate hydrogen adsorption. The adjusted electrons of the NiO/MoO2 interface are shown in Fig. 5e. The computed electron density difference shows the increased electron density (violet) of Ni atoms and decreased electron density (yellow) of Mo atoms, consistent with XPS results. Those results show that the electron distribution and energy band structures on the abundant interface of the NiO/MoO2 heterostructure can be effectively regulated, as schematically illustrated in Fig. 5f, thus coordinating the adsorption and desorption capacity of the catalyst surface to H, therefore finally improving the HER electrocatalytic activity. Combined with all the above discussions, the final NiO/MoO2-100-2 exhibits outstanding HER performance resulting from the optimized surface electron density and adsorption/desorption of intermediates as well as more available active sites.
Footnote |
† Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d2qi00136e |
This journal is © the Partner Organisations 2022 |