Al-O bridged NiFeOx/BiVO4 photoanode for exceptional photoelectrochemical water splitting

Lina Wang Hairu Wang Qian Bu Qiong Mei Junbo Zhong Bo Bai Qizhao Wang

Citation:  Lina Wang, Hairu Wang, Qian Bu, Qiong Mei, Junbo Zhong, Bo Bai, Qizhao Wang. Al-O bridged NiFeOx/BiVO4 photoanode for exceptional photoelectrochemical water splitting[J]. Chinese Chemical Letters, 2025, 36(4): 110139. doi: 10.1016/j.cclet.2024.110139 shu

Al-O bridged NiFeOx/BiVO4 photoanode for exceptional photoelectrochemical water splitting

English

  • Photoelectrochemical (PEC) water splitting has the potential to convert solar energy into sustainable hydrogen energy using semiconductor photocatalysts, thus realizing an efficient, green and carbon-free pathway for green hydrogen production [1, 2]. Narrow-band-gap metal oxides such as WO3 (~2.7 eV) [3-5], α-Fe2O3 (~2.2 eV) [5, 6] and BiVO4 (2.4 eV) [7, 8] are frequently utilized as photoanodes due to their inexpensiveness, superior photoactivity and great photostability. In particular, monoclinic phase BiVO4 is regarded as an excellent photoanode due to its non-toxicity, abundance, suitable band gap and valence band potential [9, 10]. Unfortunately, the low photogenerated electron-hole separation efficiency and slow surface water splitting rate on BiVO4 photoanode surface constrain the further improvement of solar-to-hydrogen (STH) efficiency [1, 11]. Introducing oxygen evolution co-catalysts (OECs), especially binary or multi transition metal-based cocatalysts (such as NiFe2O4/BiVO4, CoFe2O4/BiVO4 [12] and ZnCo2O4/BiVO4 [13]), helps to encourage the transfer and separation of charges, thereby enhancing charge injection efficiency and overall photocurrent density. However, the interface of OECs/BiVO4 heterojunction usually acts as a charge recombination center, resulting in lower charge transfer efficiency and terrible long-term durability. Therefore, it is of particular interest to modify the interface between OECs and BiVO4 photoanode that inhibits electron/hole pairs′ recombination and accelerating water oxidation.

    Introducing carrier shuttle bridge between OECs and BiVO4 photoanode has shown to be a workable solution to eliminate the inherent charge transfer barriers [14-16]. For example, bridging the coordination of metal hydroxides and BiVO4 photoanode with urea can achieve stable and efficient solar-driven PEC water splitting with a photocurrent density of 4.8 mA/cm2 at 1.23 V vs. RHE [17]. This is due to the fact that metal ion and urea’s strong coordination connection may prevent electron-hole pairs from recombining at the OECs/BiVO4 heterojunction, increasing the transmission of charge to the water oxidation reactive sites. However, the reported PEC water splitting performance of BiVO4-based photoanodes with the built charge shuttle bridge is far from satisfactory when compared with the theoretical photocurrent density of BiVO4 photoanode (7.5 mA/cm2) [18, 19]. Searching a more efficeint charge shuttle bridge is thus of significance and urgency. Aluminum oxide (Al2O3) is a polycrystalline structure which is composed by oxygen and aluminum atoms through ionic and covalent bonds [20-22]. Consequently, the NiFeOx/BiVO4 photoanode modified with Al-O bridge is expected to achiesubstantially improved PEC water splitting performance by virtue of the good light transmission and extremely high chemical stability [23, 24].

    Herein, we successfully bridged NiFeOx and BiVO4 with the Al-O unit through a facile impregnation method. Benefiting from the dual improved carrier charge transport by the co-loaded NiFeOx and Al2O3, this novel Al-O bridged NiFeOx/BiVO4 photoanode achieved an impressive photocurrent density of 5.87 mA/cm2 at 1.23 V vs. RHE. A stoichiometric hydrogen and oxygen evolution of 222.64 µmol/cm2 and 110.31 µmol/cm2 achieved over Al-O bridged NiFeOx/BiVO4 photoanode, which is 2.26 and 18.06 times more than NiFeOx/BiVO4 and pristine BiVO4 photoanode, respectively. This effort delivers a novel perspective to manipulate the interface of OECs/BiVO4 photoanode for highly efficient solar-driven PEC water splitting.

    The surface morphology and microstructure of the prepared films were investigated by scanning electron microscopy (SEM). From the top view (Fig. 1a and Fig. S1 in Supporting information) and the corresponding side view (Fig. S2a in Supporting information), a uniform layer of BiVO4 with irregular particles between 200 nm and 500 nm has been successfully grown uniformly on the FTO and in close contact with it. The surface and side SEM images of Al2O3/BiVO4, NiFeOx/BiVO4, NiFeOx/Al2O3/BiVO4 photoanodes are shown in Figs. 1b-d and Figs. S2b-d (Supporting information), respectively, which show that they are highly similar to the morphology of BiVO4 nanoparticles with or without the introduction of Al2O3 and NiFeOx, indicating that the morphology of the original BiVO4 nanoparticles is not changed by the simple preparation process. In order to further understand the microstructure of NiFeOx/Al2O3/BiVO4 photoanodes, a detailed study was carried out using transmission electron microscopy (TEM) and high resolution transmission electron microscope (HRTEM). The BiVO4 has a granular morphology in Fig. 1e and Fig. S3 (Supporting information) and an amorphous/disordered layer appears at the edge of BiVO4, indicating that NiFeOx and Al2O3 are coated on the surface of BiVO4 particles. Further, a clearer HRTEM image of the NiFeOx/Al2O3/BiVO4 photoanodes and the corresponding crystal plane spacing measurements were obtained in Fig. 1f by inverse Fourier transform, the spacing of the 10 ordered photoanodes was measured to be 3.1 nm, which represents the lattice stripe d = 0.31 nm matching the (121) crystal plane of BiVO4. The component mapping and energy dispersive X-Ray spectroscopy (EDX) distribution of NiFeOx/Al2O3/BiVO4 shows that Ni, Fe, Al, Bi, V and O components are evenly scattered in the nanomorphology of the sample (Figs. 1g-l and Fig. S4 in Supporting information), which further confirms that we have successfully combined NiFeOx/Al2O3/BiVO4 photoanodes by a simple process.

    Figure 1

    Figure 1.  Top-view SEM images of (a) BiVO4, (b) Al2O3/BiVO4, (c) NiFeOx/BiVO4 and (d) NiFeOx/Al2O3/BiVO4 photoanodes. HRTEM images (e, f) and TEM-EDS mapping images of NiFeOx/Al2O3/BiVO4 film (g-l).

    To further investigate the structures of BiVO4, Al2O3/BiVO4, NiFeOx/BiVO4, NiFeOx/Al2O3/BiVO4 samples, X-ray diffraction (XRD) analysis was performed in Fig. S5 (Supporting information). All samples give similar diffraction peaks as monoclinic phase scheelite BiVO4 (JCPDS No. 14–0688), appearing at 2θ of 18.6°, 18.9°, 28.9°, 30.5°, 35.2°, 47.2° and 53.1°, corresponding to (110), (011), (121), (040), (002), (042), (222) crystallographic planes of BiVO4, respectively [25]. No characteristic peaks of Al2O3 and NiFeOx were observed, probably due to the thinness of the loaded Al2O3 and NiFeOx layers or amorphous state, which is consistent with the TEM test results.

    To further reveal the surface chemistry and bonding state of the elements, X-ray photoelectron spectroscopy (XPS) analysis was performed on NiFeOx/Al2O3/BiVO4 and NiFeOx/BiVO4 photoanodes. According to Fig. S6 (Supporting information), the XPS spectrum reveals the following elements in the sample: Bi, V, O, Al, Ni and Fe. As far as XPS spectra are concerned, there are no noticeable change in position or intensity of Bi and V elements before and after the introduction of the Al2O3 interlayer in Figs. 2a and b. There are two peaks at 529.7 eV and 531.7 eV in the O 1s XPS spectra in Fig. 2c, corresponding to the binding energy of lattice oxygen and hydroxyl adsorbed oxygen [26, 27], respectively. The intensity of the lattice oxygen peak is significantly higher in NiFeOx/Al2O3/BiVO4 than in NiFeOx/BiVO4 sample, which is associated with Al2O3. Significantly, Al 2p signals were detected at 74.5 eV and 69.7 eV in Fig. 2d, which correspond to Al3+ in Al2O3 and Al3+ in tetrahedral positions [28, 29], respectively. The presence of NiFeOx is further confirmed by the observation of faint Fe and Ni signals in Figs. 2e and f. Compared with the NiFeOx/BiVO4 sample, a shift of Fe and Ni binding energy to lower binding energy was observed in the NiFeOx/Al2O3/BiVO4 photoanode, revealing an off-domain effect due to the electron shuttle bridge Al2O3 to adjust the NiFeOx/Al2O3/BiVO4 to the lowest energy [24, 29, 30]. The XPS results suggest that Al2O3 was successfully integrated into the NiFeOx/BiVO4 anode as an efficient bridge for charge transfer and further validate the successful synthesis of NiFeOx/Al2O3/BiVO4 photoanode.

    Figure 2

    Figure 2.  (a) Bi 4f, (b) V 2p, (c) O 1s, (d) Al 2p, (e) Fe 2p and (f) Ni 2p XPS spectra of NiFeOx/BiVO4 and NiFeOx/Al2O3/BiVO4.

    By using 1 mol/L potassium borate buffer (KPi, pH 9.5) and AM 1.5G (100 mW/cm2), the PEC performance of the as-prepared BiVO4 photoanodes was tested by linear scanning voltammetry (LSV). The photocurrent density of unmodified BiVO4 was 1.65 mA/cm2 at 1.23 V vs. RHE in Fig. 3a. Such low current density was ascribed by the severe charge recombination and low surface charge transport [25, 31]. The photocurrent density rose to 3.26 mA/cm2 when BiVO4 was loaded by Al2O3 (Fig. S7 in Supporting information). The photocurrent densities of the prepared NiFeOx/BiVO4 and NiFeOx/Al2O3/BiVO4 photoanodes were significantly increased after the NiFeOx co-catalyst was loaded, which is backed up by the recent literature claim that NiFeOx co-catalyst can improve BiVO4 photoanode PEC water oxidation activity [32, 33]. Differently, the prepared NiFeOx/Al2O3/BiVO4 film exhibit a significant photocurrent density of 5.87 mA/cm2 at 1.23 V vs. RHE, which is nearly 1.4 times greater than that of NiFeOx/BiVO4 (4.21 mA/cm2) and 3.5 times higher than that of pure BiVO4. The presence of the Al2O3 intermediate layer and the synergistic effect of the NiFeOx co-catalyst might be the cause of the higher photocurrent density in the NiFeOx/Al2O3/BiVO4 system. To further reveal the response of the photoanode to light under illumination with the transfer and separation of carriers, we performed transient photocurrent tests. The rapid change of photocurrent density implies a very fast charge transfer in the photoanode. It is clearly observed that the current density of NiFeOx/Al2O3/BiVO4 photoanode is the strongest and more stable after compounding NiFeOx and Al2O3 in Fig. 3b, which is the same as the previous experimental results of LSV. Assumedly, photogenerated holes on the electrode surface can be 100% injected into the electrolyte because Na2SO3 is oxidized by holes so quickly [34]. The LSV curves of the prepared BiVO4 photoanodes were tested under the conditions of 1 mol/L phosphate buffer (KPi, pH 9.5) electrolyte with the addition of Na2SO3 in Fig. 3c. The trends of the photocurrent densities were the same and increased for all photoanodes, indicating that the photoanode suffered from a severe recombination of electrons and holes for PEC water splitting performance.

    Figure 3

    Figure 3.  PEC performance of BiVO4, Al2O3/BiVO4, NiFeOx/BiVO4 and NiFeOx/Al2O3/BiVO4 photoanodes: (a) Linear-sweep voltammogram curves at scan rate of 50 mV/s. (b) Transient photocurrent (i-t) spectra at 1.23 V vs. RHE, (c) LSV curves in the presence of 1 mol/L Na2SO3 with a scan rate of 50 mV/s. H2 (d) and O2 (e) gas evolution from PEC water splitting in a 1.0 mol/L potassium borate buffer solution electrolyte. (f) Long-time i-t spectra of H2 and O2 production of NiFeOx/Al2O3/BiVO4 photoanode.

    To further demonstrate that the measured photogenerated currents were used for the PEC water splitting of hydrogen and oxygen production, the hydrogen and oxygen produced by BiVO4, Al2O3/BiVO4, NiFeOx/BiVO4 and NiFeOx/Al2O3/BiVO4 photoanodes were quantified using gas chromatography. The Figs. 3d and e demonstrate that the gas precipitation from all photoanodes showed an increasing trend with time. After 120 min of continuous light exposure to the PEC decomposition water reaction, the yields of H2 and O2 from BiVO4, Al2O3/BiVO4, NiFeOx/BiVO4 and NiFeOx/Al2O3/BiVO4 photoanodes were 12.33 and 5.66 µmol/cm2, 72.14 and 35.06 µmol/cm2, 98.42 and 49.07 µmol/cm2 and 222.64 and 110.31 µmol/cm2, respectively (Fig. S9 in Supporting information). The NiFeOx/Al2O3/BiVO4 photoanode for PEC decomposition of water performance was nearly 18 times that of the blank BiVO4, which verified the feasibility of our design using Al2O3 as an electron shuttle bridge and the introduction of NiFeOx co-catalyst for improving BiVO4′s PEC performance. In the meanwhile, the photoelectrodes’ stability test revealed that after 120 min of exposure to light at 1.23 V vs. RHE, the photocurrent density of the NiFeOx/Al2O3/BiVO4 electrode was greater than that of BiVO4, indicating a significant enhancement of the photostability of NiFeOx/Al2O3/BiVO4 (Fig. 3f). Meanwhile, the XRD pattern and SEM in Fig. S10 (Supporting information) showed little change in the NiFeOx/Al2O3/BiVO4 photoanode morphology before and after the test, further confirming the enhanced stability of the BiVO4 photoanode after deposition of NiFeOx and Al2O3. Therefore, the NiFeOx/Al2O3/BiVO4 thin film not only can avoid the oxidation of BiVO4, but also can promote the transfer of carriers to the electrolyte solution, thus promoting the release of hydrogen and oxygen.

    UV–vis absorption spectra were used to check the synthetic photoanodes’ capacity to absorb light. In accordance with other results [18, 31, Fig. 4a demonstrates that BiVO4 exhibits a considerable light absorption at a wavelength of around 500 nm. The light absorption of BiVO4 photoanode has been slightly broadened after the introduction of Al2O3 and NiFeOx. Correspondingly, the optical bandgaps of BiVO4, Al2O3/BiVO4, NiFeOx/BiVO4 and NiFeOx/Al2O3/BiVO4 calculated by Tauc method are 2.47 eV, 2.46 eV, 2.46 eV and 2.43 eV, respectively (Fig. S11 in Supporting information). As a result, the NiFeOx/Al2O3/BiVO4 photoanode’s improvements in PEC water splitting activity are not primarily influenced by the energy band structure of the photoanodes. PEC activity and incident light wavelength can be related to one other using observations of incident photon to current conversion efficiency (IPCE) in Fig. 4b. Pure BiVO4 shows the lowest IPCE value in the 300–550 nm. However, the IPCE value increased and reached 45% at 430 nm for Al2O3/BiVO4. Similarly, the IPCE values of NiFeOx/BiVO4 and NiFeOx/Al2O3/BiVO4 show a further increase. In particular, the NiFeOx/Al2O3/BiVO4 shows the largest IPCE value of 58% at 430 nm, demonstrating that NiFeOx and Al2O3 do not expand the light absorption, but can significantly increase the efficiency of solar energy conversion and the interfacial separation of photogenerated carriers. The applied bias photon-to-current efficiency (ABPE) was computed from the LSV curve to describe the solar water splitting half-reaction hydrogen precipitation efficiency. Accordingly, the NiFeOx/Al2O3/BiVO4 photoanode has a higher photoelectric conversion efficiency in the whole applied bias range and attains an ABPE of 2.19% at 0.67 V vs. RHE in Fig. 4c, further supporting the claim that combining NiFeOx and Al2O3 is a successful method for enhancing BiVO4’s PEC performance with higher photoelectric conversion efficiency at lower potentials. In comparison, this was found to be an extremely high level among the many photoanodes reported so far (Fig. 4d) [35-38].

    Figure 4

    Figure 4.  The Optical properties and PEC properties of BiVO4, Al2O3/BiVO4, NiFeOx/BiVO4 and NiFeOx/Al2O3/BiVO4 photoanodes: (a) the UV–vis absorption spectra; (b) IPCE curves at 1.23 V vs. RHE; (c) ABPE curves and (d) the ABPE of different photoelectrodes under AM1.5G illumination.

    Based on the aforementioned experimental results, the values of surface charge injection efficiency (ηinj) and bulk phase charge separation efficiency (ηsep) of the samples were determined. Compared with the pure BiVO4 photoanode, the ηinj of Al2O3/BiVO4, NiFeOx/BiVO4 and NiFeOx/Al2O3/BiVO4 have significantly improved in the overall bias voltage range in Fig. S12a (Supporting information), especially for the NiFeOx/BiVO4 electrode. The ηinj of NiFeOx/BiVO4 shows a decreasing trend when the bias voltage is > 1.3 V vs. RHE, indicating that the water oxidation reaction at the interface has almost reached saturation at this time [39], further demonstrating the effectiveness of NiFeOx in accelerating the water oxidation reaction kinetics. The ηsep follows a similar pattern with ηsep reaching 74.4% for NiFeOx/BiVO4 and 62.4% for Al2O3/BiVO4 at 1.23 V vs. RHE in Fig. S12b (Supporting information), both of which are much greater than BiVO4 (59.3%). After introducing Al2O3 between NiFeOx/BiVO4, the ηsep of NiFeOx/Al2O3/BiVO4 is increased to 85.2% at 1.23 V vs. RHE, indicating that Al2O3 forms a fast transport channel for photogenerated electrons, reducing their chance of recombination in the bulk phase. Electrochemical impedance spectroscopy (EIS) measurements were carried out in Fig. S12c (Supporting information) under visible light irradiation to acquire a better understanding of the electron transfer kinetics. From the impedance diagrams of Fig. S12c and simulated impedance values of Table S1 (Supporting information), the impedances of Al2O3/BiVO4, NiFeOx/BiVO4 and NiFeOx/Al2O3/BiVO4 photoanodes are all reduced compared to the original BiVO4 photoanode, suggesting that the addition of both Al2O3 and NiFeOx helps to improve the electrical conductivity of the BiVO4 photoanode while the addition of Al2O3 optimizes the contact between NiFeOx and BiVO4 substrate, which may be one of the reasons for its high photocurrent density. To demonstrate the charge transport mechanism of BiVO4, Al2O3/BiVO4, NiFeOx/BiVO4 and NiFeOx/Al2O3/BiVO4 photoanodes, the room temperature steady-state PL were characterized in Fig. S12d (Supporting information). As a consequence, BiVO4 showed a signal peak in the wavelength range around 500 nm, which may be due to the emission of the BiVO4 band gap jump. Likewise, the signal peaks of Al2O3/BiVO4, NiFeOx/BiVO4 and NiFeOx/Al2O3/BiVO4 showed no change in position but the intensity of the emission peak was significantly reduced, indicating that the NiFeOx/Al2O3/BiVO4 photoanode could prevent the photogenerated electron-hole pair recombination.

    Mott-Schottky (M-S) experiments were performed in Fig. 5a to further study the intrinsic electronic structure of the photoanodes. The positive slopes of the M-S curves show that these photoanodes are n-type semiconductors with a high proportion of migrating electrons. The carrier concentrations of BiVO4, Al2O3/BiVO4, NiFeOx/BiVO4 and NiFeOx/Al2O3/BiVO4 photoanodes were calculated to be 2.91 × 1021, 3.12 × 1021, 3.21 × 1021 and 3.97 × 1021 cm−3 in Table S2 (Supporting information), respectively. The carrier concentration did not change significantly and also verified that the Al, Ni and Fe ions were not doped into the BiVO4 photoanode during the loading of Al2O3 and NiFeOx. The OCP curves were obtained by rapidly switching off the light source after reaching steady state at the open circuit voltage to analyze the mechanism of charge separation. In Fig. 5b, two decay process are observed after switching off the light. The instantaneous rise in voltage indicates that part of the electrons has returned to the semiconductor’s valence band (VB), while the other part is captured and recombinated with holes. The later, slowly decaying process represents the injection of charge into the electrolyte. Compared to the other samples, the NiFeOx/Al2O3/BiVO4 photoanode has a more positive potential value of 0.21 V vs. RHE, while the BiVO4 film has a potential value of 0.38 V vs. RHE, further demonstrating that the co-addition of NiFeOx and Al2O3 can accelerate the aggregation of photogenerated electrons at the conduction band (CB) of BiVO4, which in turn induces the NiFeOx/Al2O3/BiVO4 surface electron-hole pair separation.

    Figure 5

    Figure 5.  Carrier transport performance of BiVO4, Al2O3/BiVO4, NiFeOx/BiVO4 and NiFeOx/Al2O3/BiVO4 photoanodes: (a) Mott-Schottky curves under dark; (b) Open circuit potential measurements.

    The energy band structure of the BiVO4 and NiFeOx/BiVO4 samples was investigated to study and clarify the reaction mechanism for PEC water splitting. In the M-S curve, the intersection of the tangent line with the x-axis (y = 0) is the flat-band potential. The flat-band potential is 0.2 V higher than the CB potential for n-type semiconductors, so the CBs of BiVO4 and NiFeOx/BiVO4 photoanodes are −0.06 V vs. NHE and −0.17 V vs. NHE (ERHE = ENHE + pH × 0.059). After combining with the previously obtained data on the forbidden band width of the films, a schematic energy band structure of the as-prepared films is obtained. From the Fig. 6, it can be found that the CB and VB of BiVO4 films are gradually shifted to lower energy levels with the addition of NiFeOx. Such a structure not only allows the excited holes of VB of BiVO4 to be transferred to the aqueous solution, but also keeps the photogenerated electrons in BiVO4’s CB.

    Figure 6

    Figure 6.  The mechanism of NiFeOx/Al2O3/BiVO4 photoanode for PEC water splitting.

    Based on the above experimental results, we propose the mechanism of charge separation and transfer at NiFeOx/Al2O3/BiVO4 electrodes in the PEC water splitting process (Fig. 6). The energy band structure of NiFeOx was adopted from previous reports [40, 41]. The electrons stored in the BiVO4 photoanode are also rapidly released and transmitted to the Pt photocathode, where they reduce water to produce H2. The holes remaining in the VB of the BiVO4 photoanode will be rapidly transferred to the NiFeOx surface via the Al-O bridge, then transferred to the electrolyte solution for the oxidation reaction. At the same time, the Al-O bridge acts as a channel to speed up the transmission of holes from the BiVO4 photoanode to the NiFeOx/Al2O3/BiVO4 photoanode surface. Therefore, the excellent PEC water splitting performance of NiFeOx/Al2O3/BiVO4 photoanode depends on the dual improvement of carrier transfer both within the photoanode and between the photoanode and electrolyte when NiFeOx and Al2O3 are co-loaded.

    To address the shortcomings of BiVO4 photoanode that electrons and holes are easily compounded, NiFeOx/Al2O3/BiVO4 photoanode was successfully designed and prepared by compounding the interlayer Al2O3, which significantly improved the PEC performance of BiVO4. The NiFeOx/Al2O3/BiVO4 photoanode not only maintains high chemical stability, but also achieves a water oxidation photocurrent density of 5.87 mA/cm2 (1.23 V vs. RHE) and an ABPE value of 2.19% (0.67 V vs. RHE), respectively, a 3.5-fold improvement over BiVO4 (1.65 mA/cm2) and 8.1 times better than BiVO4 (0.27%). By modifying the Al2O3 layer with a NiFeOx co-catalyst, the holes are rapidly transferred to the NiFeOx surface for the water oxidation reaction, which reduces the interfacial charge recombination, regulates carrier concentration, effectively promotes charge separation and improves the stability of PEC water splitting. This research offers fresh perspectives and takeaways for the practical use of solar-powered PEC water splitting systems.

    All authors declared that they do not have any commercial or associative interest that represents a conflict of interest in connection with the work submitted.

    Lina Wang: Writing – original draft, Writing – review & editing. Hairu Wang: Data curation, Writing – review & editing. Qian Bu: Writing – review & editing. Qiong Mei: Data curation, Methodology, Writing – review & editing. Junbo Zhong: Supervision, Writing – review & editing. Bo Bai: Resources, Supervision. Qizhao Wang: Funding acquisition, Resources, Supervision, Writing – review & editing.

    This work was financially supported by the National Natural Science Foundation of China (No. 52173277), the Fundamental Research Funds for the Central Universities of Chang’an University (No. 300102299304), the Innovative Research Team for Science and Technology of Shaanxi Province (No. 2022TD-04) and the open program of Key Laboratories of Fine Chemicals and Surfactants in Sichuan Provincial Universities (No. 2023JXZ03).

    Supplementary material associated with this article can be found, in the online version, at doi:10.1016/j.cclet.2024.110139.


    1. [1]

      X. Meng, C. Zhu, X. Wang, et al., Nat. Commun. 14 (2023) 2643. doi: 10.1038/s41467-023-38138-9

    2. [2]

      M. Liu, G. Zhang, X. Liang, et al., Angew. Chem. Int. Ed. 135 (2023) e202304694. doi: 10.1002/ange.202304694

    3. [3]

      W. Zhang, M. Tian, H. Jiao, et al., Chin. J. Catal. 43 (2022) 2321–2331.

    4. [4]

      H. Sun, W. Hua, Y. Li, et al., ACS Sustain. Chem. Eng. 8 (2020) 12637–12645. doi: 10.1021/acssuschemeng.0c04204

    5. [5]

      F. Xiao, R. Guo, X. He, et al., Int. J. Hydrogen Energy 46 (2021) 7954–7963.

    6. [6]

      N. Duc Quang, P. Cao Van, S. Majumder, et al., J. Colloid Interface Sci. 616 (2022) 749–758.

    7. [7]

      V. Andrei, G.M. Ucoski, C. Pornrungroj, et al., Nature 608 (2022) 518–522. doi: 10.1038/s41586-022-04978-6

    8. [8]

      H. Bai, F. Wang, Z. You, et al., Colloids Surf. A 640 (2022) 128412. doi: 10.1016/j.colsurfa.2022.128412

    9. [9]

      J. Zhang, X. Wei, J. Zhao, et al., Chem. Eng. J. 454 (2023) 140081. doi: 10.1016/j.cej.2022.140081

    10. [10]

      W. Zhou, T. Jiang, Y. Zhao, et al., J. Colloid Interface Sci. 549 (2019) 42–49.

    11. [11]

      W. Bai, Y. Zhou, G. Peng, et al., Appl. Catal. B 315 (2022) 121606. doi: 10.1016/j.apcatb.2022.121606

    12. [12]

      Q. Wang, J. He, Y. Shi, et al., Appl. Catal. B 214 (2017) 158–167.

    13. [13]

      J. Huang, Y. Wang, K. Chen, et al., Chin. Chem. Lett. 33 (2022) 2060–2064. doi: 10.1016/j.cclet.2021.08.082

    14. [14]

      Y. Dou, X. Yang, Q. Wang, et al., J. Colloid Interface Sci. 644 (2023) 256–263. doi: 10.1016/j.jcis.2023.04.082

    15. [15]

      X. Wang, H. Jiang, M. Zhu, et al., Chin. Chem. Lett. 34 (2023) 107683. doi: 10.1016/j.cclet.2022.07.026

    16. [16]

      X. Xu, H. Wang, K. Chen, et al., Chem. Eng. J. 485 (2024) 149922. doi: 10.1016/j.cej.2024.149922

    17. [17]

      M. Sun, C. Yuan, R.T. Gao, et al., Chem. Eng. J. 426 (2021) 131062. doi: 10.1016/j.cej.2021.131062

    18. [18]

      Y. Li, Q. Wang, X. Hu, et al., Chem. Eng. J. 433 (2022) 133592. doi: 10.1016/j.cej.2021.133592

    19. [19]

      Y. Long, H. Xu, J. He, et al., Surf. Interfaces 31 (2022) 102056. doi: 10.1016/j.surfin.2022.102056

    20. [20]

      Z. Liu, G. Xu, L. Zeng, et al., Appl. Catal. B 324 (2023) 122259. doi: 10.1016/j.apcatb.2022.122259

    21. [21]

      Y. Zhang, M. Choi, Z. Wang, et al., Appl. Surf. Sci. 609 (2023) 155295. doi: 10.1016/j.apsusc.2022.155295

    22. [22]

      D.H.K. Jackson, T.F. Kuech, J. Power Sources 365 (2017) 61–67. doi: 10.1016/j.jpowsour.2017.08.076

    23. [23]

      R. Rajendiran, P.K. Seelam, A. Patchaiyappan, et al., Chem. Eng. J. 451 (2023) 138507. doi: 10.1016/j.cej.2022.138507

    24. [24]

      Y. Zhao, G. Liu, H. Wang, et al., J. Mater. Chem. A 9 (2021) 11285–11290. doi: 10.1039/d1ta00206f

    25. [25]

      L. Wang, Z. Liu, J. Zhang, et al., Chin. Chem. Lett. 34 (2023) 108007. doi: 10.1016/j.cclet.2022.108007

    26. [26]

      W. Chen, J. Du, H. Zhang, et al., Chin. Chem. Lett. (2023), doi: https://doi.org/ 10.1016/j.cclet.2023.109168.

    27. [27]

      H. Ming, X. Bian, J. Cheng, et al., Appl. Catal. B 341 (2024) 123314. doi: 10.1016/j.apcatb.2023.123314

    28. [28]

      Z. Boukha, J.L. Ayastuy, A. Iglesias-González, et al., Appl. Catal. B 160-161 (2014) 629–640. doi: 10.1016/j.apcatb.2014.06.002

    29. [29]

      Z. Zhou, S. Wu, L. Li, et al., ACS Appl. Mater. Interfaces 11 (2019) 5978–5988. doi: 10.1021/acsami.8b18681

    30. [30]

      J. Wang, C. Qin, H. Wang, et al., Appl. Catal. B 221 (2018) 459–466. doi: 10.1016/j.apcatb.2017.09.042

    31. [31]

      K. Chen, R. Wang, Q. Mei, et al., Appl. Catal. B 344 (2024) 123670. doi: 10.1016/j.apcatb.2023.123670

    32. [32]

      L. Ling, C. Yuan, Q. Xu, et al., Surf. Interfaces 36 (2023) 102483. doi: 10.1016/j.surfin.2022.102483

    33. [33]

      S. Wang, T. He, P. Chen, et al., Adv. Mater. 32 (2020) 2001385. doi: 10.1002/adma.202001385

    34. [34]

      L. Wang, H. Cheng, Z. Zhang, et al., Chem. Eng. J. 456 (2023) 140990. doi: 10.1016/j.cej.2022.140990

    35. [35]

      C. Feng, Q. Zhou, B. Zheng, et al., J. Mater. Chem. A 7 (2019) 22274–22278. doi: 10.1039/c9ta08951a

    36. [36]

      X. Ning, B. Lu, Z. Zhang, et al., Angew. Chem., Int. Ed. 58 (2019) 16800–16805. doi: 10.1002/anie.201908833

    37. [37]

      H. Chen, S. Wang, J. Wu, et al., J. Mater. Chem. A 8 (2020) 13231–13240. doi: 10.1039/d0ta04572a

    38. [38]

      S. Wang, P. Chen, Y. Bai, et al., Adv. Mater. 30 (2018) 1800486. doi: 10.1002/adma.201800486

    39. [39]

      Y. Wang, J. Huang, L. Wang, et al., Chin. J. Struct. Chem. 41 (2022) 2201054–2201068.

    40. [40]

      S. Seenivasan, S. Adhikari, D.H. Kim, Chem. Eng. J. 422 (2021) 130137.

    41. [41]

      F. Ning, M. Shao, S. Xu, et al., Energy Environ. Sci. 9 (2016) 2633–2643.

  • Figure 1  Top-view SEM images of (a) BiVO4, (b) Al2O3/BiVO4, (c) NiFeOx/BiVO4 and (d) NiFeOx/Al2O3/BiVO4 photoanodes. HRTEM images (e, f) and TEM-EDS mapping images of NiFeOx/Al2O3/BiVO4 film (g-l).

    Figure 2  (a) Bi 4f, (b) V 2p, (c) O 1s, (d) Al 2p, (e) Fe 2p and (f) Ni 2p XPS spectra of NiFeOx/BiVO4 and NiFeOx/Al2O3/BiVO4.

    Figure 3  PEC performance of BiVO4, Al2O3/BiVO4, NiFeOx/BiVO4 and NiFeOx/Al2O3/BiVO4 photoanodes: (a) Linear-sweep voltammogram curves at scan rate of 50 mV/s. (b) Transient photocurrent (i-t) spectra at 1.23 V vs. RHE, (c) LSV curves in the presence of 1 mol/L Na2SO3 with a scan rate of 50 mV/s. H2 (d) and O2 (e) gas evolution from PEC water splitting in a 1.0 mol/L potassium borate buffer solution electrolyte. (f) Long-time i-t spectra of H2 and O2 production of NiFeOx/Al2O3/BiVO4 photoanode.

    Figure 4  The Optical properties and PEC properties of BiVO4, Al2O3/BiVO4, NiFeOx/BiVO4 and NiFeOx/Al2O3/BiVO4 photoanodes: (a) the UV–vis absorption spectra; (b) IPCE curves at 1.23 V vs. RHE; (c) ABPE curves and (d) the ABPE of different photoelectrodes under AM1.5G illumination.

    Figure 5  Carrier transport performance of BiVO4, Al2O3/BiVO4, NiFeOx/BiVO4 and NiFeOx/Al2O3/BiVO4 photoanodes: (a) Mott-Schottky curves under dark; (b) Open circuit potential measurements.

    Figure 6  The mechanism of NiFeOx/Al2O3/BiVO4 photoanode for PEC water splitting.

  • 加载中
计量
  • PDF下载量:  2
  • 文章访问数:  93
  • HTML全文浏览量:  6
文章相关
  • 发布日期:  2025-04-15
  • 收稿日期:  2024-02-14
  • 接受日期:  2024-06-17
  • 修回日期:  2024-05-19
  • 网络出版日期:  2024-06-18
通讯作者: 陈斌, bchen63@163.com
  • 1. 

    沈阳化工大学材料科学与工程学院 沈阳 110142

  1. 本站搜索
  2. 百度学术搜索
  3. 万方数据库搜索
  4. CNKI搜索

/

返回文章