Yang
Li
a,
Shumei
Chen
*a,
Xin
Wu
b,
Huabin
Zhang
*c and
Jian
Zhang
b
aCollege of Chemistry, Fuzhou University, Fuzhou 350108, China. E-mail: csm@fzu.edu.cn
bState Key Laboratory of Structural Chemistry, Fujian Institute of Research on the Structure of Matter, Chinese Academy of Sciences, Fuzhou, 350002, China
cKAUST Catalysis Center (KCC), Physical Sciences and Engineering Division, King Abdullah Uniersity of Science and Technology (KAUST), Thuwal 23955-6900, Saudi Arabia. E-mail: Huabin.Zhang@kaust.edu.sa
First published on 23rd July 2021
Sustainable hydrogen fuel supply through electrochemical water splitting requires highly efficient, low-cost and robust electrocatalysts. Interface engineering is of key importance to improve the catalytic performance in a heterogeneous electrocatalytic system. Herein, a porous microcubic framework composed of a ZnO/ZnMoO4 heterostructure (ZnO@ZnMoO4) is prepared by a hybrid zeolitic imidazolate framework-derived oxidation method, and it shows much enhanced hydrogen evolution reaction (HER) activity in alkaline media. The overpotential (at 10 mA cm−2) for ZnO@ZnMoO4 is significantly reduced by 30% and 20% compared with those for virgin ZnO (v-ZnO) and polycrystalline zinc molybdenum oxide (PZMO), respectively. The enhanced electrocatalytic activity should be attributed to the ZnO/ZnMoO4 heterostructure, which can synergistically facilitate the charge transport. This work provides a more structured design strategy for electrocatalysts for future electrochemical energy conversion systems.
The low intrinsic conductivity of MBO electrocatalysts limits their widespread use.17 Construction of heterostructured electrocatalysts can serve as an efficacious approach to achieve high performance in the electrocatalytic HER.18–20 Compared with a single catalyst, heterostructured electrocatalysts could show a combination of the strengths of their parent phases.21,22 Moreover, the built-in electric fields or chemical potential differences at the heterostructure can provide a driving force for charge transfer.23–25 Co-aggregation and epitaxial growth are two common methods for the preparation of heterostructured electrocatalysts, but tend to lower their accessible active surface area.26 Therefore, it is crucial to explore new methods for realizing the nano-sized heterostructure of MBO electrocatalysts.
Metal–organic frameworks (MOFs) have emerged as an interesting class of functional materials and precursors of inorganic materials for electrochemical energy-related applications.27–30 As a subfamily of MOFs, Zn, Mo-based hybrid zeolitic imidazolate frameworks (HZIF-Zn/Mo) with excellent stability and porous structures demonstrate great potential in the electrocatalytic field.31–33 The atomic-level structural uniformity of HZIFs can realize the intergrowth of different functional materials at the molecular level under desired conditions, resulting in the construction of a Mo-based heterostructured catalyst in a flexible and controllable manner.34 Zhang et al. have constructed an ultrasmall MoC/MoS2 heterostructure using HZIFs as the precursor and the obtained MoCSx heterostructure exhibits much enhanced oxygen reduction activity.35 In this regard, the HZIF-assisted strategy can be an economical strategy for preparing heterostructured MBO electrocatalysts.
Herein, we prepared ZnO@ZnMoO4via direct oxidation of HZIFs and systematically evaluated its HER performance in an alkaline medium. The obtained ZnO@ZnMoO4 integrates the structural advantages of different functional subunits. Specifically, the variable oxidation states of the elements in ZnMoO4 ensure more active sites for the HER. The electrical coupling at the ZnO/ZnMoO4 heterostructure is beneficial for electron transfer and structural stability. Moreover, benefiting from the mutual promoted action at the interface by the synergistic effect of two components, reconstructed active centers with higher intrinsic activity are established. Impressively, the obtained ZnO@ZnMoO4 hybrid exhibits much boosted HER activity in alkaline media.
Post-annealing of the performed HZIF-Zn/Mo under air results in the formation of ZnO@ZnMoO4. In this process, the organic moieties in HZIF-Zn/Mo are evaporated at a high temperature, whereas the Zn and Mo components react with O to form ZnO and ZnMoO4. The X-ray diffraction (XRD) spectrum shows the coexistence of ZnO and ZnMoO4 in ZnO@ZnMoO4 (Fig. S1†).36 Scanning electron microscopy (SEM) and transmission electron microscopy (TEM) images show that the obtained ZnO@ZnMoO4 retains the original cubic shape from the parent HZIF-Zn/Mo and exhibits a rougher surface (Fig. 1a–e and S2†). A closer observation of the cracked particle reveals that the interior of ZnO@ZnMoO4 is composed of numerous loosely assembled nanoparticles (Fig. 1b and c). The formation of the loose structure probably originates from the consumption of the organic moiety in the pyrolysis process. The high-resolution TEM (HRTEM) image shows clear lattice fringes with lattice spacings of 0.28 and 0.23 nm, corresponding to the (100) and (130) crystal planes of ZnO and ZnMoO4, respectively, revealing the formation of an intimate ZnO/ZnMoO4 heterostructure (Fig. 1f).37,38 Energy-dispersive X-ray spectroscopy (EDS) investigation confirms that ZnO@ZnMoO4 is composed of Zn, Mo and O (Fig. S3†). High-angle annular dark-field scanning transmission electron microscopy (HAADF-STEM) coupled with elemental mapping images show the homogeneous distribution of O, Zn and Mo in the ZnO@ZnMoO4 framework (Fig. 1g–i). Virgin ZnO (v-ZnO) was fabricated by performing a similar synthetic process with ZnO@ZnMoO4 except for using ZIF-8 as the precursor and polycrystalline zinc molybdenum oxide (PZMO) was also prepared and they were used as the references (Figs. S4 and S5†).
![]() | ||
Fig. 1 (a–c) SEM images of ZnO@ZnMoO4. (d and e) TEM images of ZnO@ZnMoO4. (f) HRTEM image of ZnO@ZnMoO4. (g–i) HAADF-STEM and elemental mapping images of ZnO@ZnMoO4. |
The surface chemical state of ZnO@ZnMoO4 is investigated by X-ray photoelectron spectroscopy (XPS). Compared with v-ZnO, additional Mo 3d, Mo 3p3/2 and Mo 3p1/2 profiles are observed for ZnO@ZnMoO4, suggesting the co-existence of Zn, Mo and O (Fig. 2a). The O 1s spectrum can be well divided into three peaks (Fig. 2b). The peaks at 529.8 and 530.8 eV correspond to the Mo–O and Zn–O bonds, respectively.39 The higher energy peak at 532 eV is ascribed to chemisorbed oxygen.40–42 In Fig. 2c, the Zn 2p3/2 and Zn 2p1/2 peaks shift toward the higher binding energy after ZnMoO4 modification, indicating the interaction of ZnO and ZnMoO4.43 In the high-resolution Mo 3d spectrum, the doublet located at the higher binding energy is attributed to the Mo6+ species, while the doublet with the lower binding energy originates from Moδ+ (δ < 6) (Fig. 2d).44 The appearance of Moδ+ could be due to the electronic redistribution at the ZnO/ZnMoO4 nanointerface via charge transfer, strongly suggesting the chemical binding of the in situ formed ZnO and ZnMoO4.45 The structural information of ZnO@ZnMoO4 is further analyzed using the Raman spectrum (RM). The E2L-E2H mode at 332 cm−1 signifies the second order non-polar scattering arising from the ZnO zone boundary centre (Fig. 2e).46 The peak at 437 cm−1 is the E2(high) mode of ZnO owing to the vibration of the oxygen atom, indicating the high crystal quality of ZnO. The disappearance of E2(high) for ZnO@ZnMoO4 demonstrates the decreased c-axis growth and worse crystallization of the ZnO component, which could be attributed to the chemical bonding of ZnO with ZnMoO4 because of their intergrowth during the annealing process.47 The E1(LO) mode observed at 583 cm−1 means the existence of the oxygen vacancy and Zn interstitial in both v-ZnO and ZnO@ZnMoO4. The N2 adsorption and desorption isotherms show that the Brunauer–Emmett–Teller (BET) surface area of ZnO@ZnMoO4 is 19.6 m2 g−1, higher than that of v-ZnO (15.5 m2 g−1) (Fig. 2f).
The HER activities of the as-prepared catalysts are investigated in 1.0 M KOH. Fig. 3a shows the linear sweep voltammetry (LSV) curves of different electrode materials. Unsurprisingly, a commercial Pt/C (20 wt%) catalyst demonstrates the best HER activity, showing near ∼0 V onset potential. v-ZnO exhibits inferior HER activity, requiring an overpotential (η10) of 0.52 V at a current density of 10 mA cm−2. A positively shifted η10 is observed for PZMO (0.45 V), indicating that more active sites existed in the constructed bimetallic oxide. The HER activity is further accelerated by the ZnO@ZnMoO4 catalyst, the η10 value of which is reduced to 0.36 V, 30%, 20% and 18% lower than those of v-ZnO, PZMO and their mechanical mixture (ZnO/PZMO), respectively. These results indicate that the interface optimizing between ZnO and ZnMoO4 can induce a synergistically enhanced HER activity.
![]() | ||
Fig. 3 (a–c) LSV curves, Tafel plots and EIS of various electrodes in 1.0 M KOH. (d) The capacitive currents as a function of scan rate for various electrodes. |
The alkaline water-splitting kinetics of various catalysts is revealed by Tafel plots derived from the polarization curves. The Tafel slope of ZnO@ZnMoO4 is confirmed to be 63 mV dec−1 and smaller than those of PZMO (85 mV dec−1), v-ZnO (479 mV dec−1) and ZnO/PZMO (72 mV dec−1), indicating the higher HER rate of ZnO@ZnMoO4 (Fig. 3b). This result suggests that the HER over ZnO@ZnMoO4 proceeds via a Heyrovsky-dominated Volmer–Heyrovsky mechanism, which indicates that the sluggish dissociation of H2O is the rate determining step.48,49 The electrode kinetics of ZnO@ZnMoO4 in the HER process is evaluated by electrochemical impedance spectroscopy (EIS). The semicircle in the low-frequency range on the Nyquist plot derives the charge-transfer resistance (Rct), which is highly related to the electrocatalytic kinetics. The EIS results show that ZnO@ZnMoO4 has a smaller Rct value than the other samples, indicating a faster charge transfer rate at the ZnO@ZnMoO4/electrolyte interface (Fig. 3c).
The boosted HER activity of ZnO@ZnMoO4 could originate from the high electrochemically active surface area (ECSA). To confirm this point, we investigate the double-layer capacitance (Cdl) which is positively proportional to the ECSA by performing cyclic voltammetry (CV) measurements. The current response in the CV test is ascribed to the exposed surface area of the electrode in the electrolyte. As shown in Fig. S6,† the highest Cdl value of ZnO@ZnMoO4 indicates the most electrolyte-accessible surface area, which could originate from its interior porous structure. The chronoamperometric response at a fixed potential of −0.4 V is measured to investigate the stability of ZnO@ZnMoO4 (Fig. S7†). During the testing, the minor fluctuations indicate the hydrogen gas generation, accumulation and desorption. A less than 20% degradation of the initial current after 40000 s is observed, demonstrating the strong durability of ZnO@ZnMoO4. The XRD, HRTEM and EDS characterization of ZnO@ZnMoO4 after the HER reveals the decreased crystallinity of the ZnMoO4 moiety, which may result from the etch in the alkaline medium during the HER process.
In summary, in this work we developed an effective strategy to accelerate the water-splitting activity of ZnMoO4 by cooperating with ZnO to form an interfacial heterostructure. The experimental results demonstrate that the boosted HER activity derives from the modulated electronic states at the heterointerface, which offer abundant active sites and improve the conductivity of ZnMoO4. Profiting from the aforementioned point, ZnO@ZnMoO4 shows a low onset overpotential, small Tafel slope and good cycling stability. This work not only develops a low-cost and efficient HER electrocatalyst, but also delivers a HZIF-derived method, which could be extended to the synthesis of other multifunctional Mo-based catalysts for energy conversion and storage applications.
Footnote |
† Electronic supplementary information (ESI) available: Experimental details, CV, and Cdl plots. See DOI: 10.1039/d1dt01861b |
This journal is © The Royal Society of Chemistry 2021 |