Exploring Bi4V2O11 as Photoanode for Water Splitting with a Wide Range of Solar Light Capture and Suitable Band Potential

Article Open Access

Exploring Bi4V2O11 as Photoanode for Water Splitting with a Wide Range of Solar Light Capture and Suitable Band Potential

Author Information
1
School of Materials Science and Engineering, Nanyang Technological University, 50 Nanyang Avenue, Singapore 639798, Singapore
2
College of Resources and Environmental Sciences, China Agricultural University, Beijing 100193, China
3
Eco-materials and Renewable Energy Research Center (ERERC), College of Engineering and Applied Science, Nanjing University, Nanjing 210093, China
4
Energy Foundation China, No. 19 Jianguomenwai Dajie, Beijing 100004, China
5
School of Chemical Engineering, Northwest University, Xi’an 710069, China
*
Authors to whom correspondence should be addressed.
Views:796
Downloads:125
Photocatalysis: Research and Potential 2024, 1 (2), 10002;  https://doi.org/10.35534/prp.2024.10002

Received: 21 February 2024 Accepted: 08 April 2024 Published: 11 April 2024

Creative Commons

© 2024 by the authors; licensee SCIEPublish, SCISCAN co. Ltd. This article is an open access article distributed under the CC BY license (https://creativecommons.org/licenses/by/4.0/).

ABSTRACT: Bi4V2O11 possesses a bandgap of ~1.9 eV, and the band positions of minimum conduction band and maximum valence band straddle the redox potentials of H+/H2 and O2/H2O. In the current work, photoanode made of particulate Bi4V2O11 film displays a wide range of light adsorption. However, when the anode was fabricated by drop-casting and examined for photoelectrochemical water splitting, the photocurrent density of the pristine Bi4V2O11 was low. Improvement has then been carried out by Mo-doping. The Mo-doped Bi4V2O11 photoanode achieves a maximum photocurrent density of 0.3 mA/cm2 after a post deposition necking treatment to improve the connectivity of the drop-cast particles in the film. This material also shows a stability with maintaining 80% photocurrent after 2 h test. Discussion has been made on the displayed performance in PEC water splitting of the Bi4V2O11 materials. Potential solutions have been proposed for this type of promising photoanode material for water splitting.
Keywords: Visible light photocatalyst; Bi4V2O11; Photoelectrochemical water splitting; Photoanode

1. Introduction

Photoelectrochemical (PEC) splitting of water into hydrogen and oxygen is a promising route to produce clean and renewable H2 energy [1,2,3]. Compared with the efficient p-type photocathodes, the low efficiency in oxygen evolution poses a great challenge on water splitting [4]. Recently, (oxy)nitride materials such as Ta3N5 [5,6,7], LaTiO2N [8,9], and SrTaO2N [10], BaTaO2N [11] have attracted a great deal of interest due to their high theoretical solar-to-hydrogen efficiency (over 15%), as well as suitable conduction and valence band positions for water splitting. However, most n-type non-oxide semiconductors are chemically unstable and easily oxidized by photo-generated holes [12]. Compared with the oxynitride materials, metal oxides are more promising to be stable photoanodes. Unfortunately, most of the currently reported metal oxides cannot harvest enough visible light due to their wide bandgaps, such as TiO2, WO3 and BiVO4 [12,13,14,15]. Although hematite (α-Fe2O3) is a narrow band gap oxide photoanode material with bandgap around 2.0 eV, the bottoms of the conduction band of hematite is more positive (around 0.5 eV) than the water reduction potential. This would require a large external bias to overcome the potential difference [16]. Besides, very short hole diffusion length (only a few nano meters) and short carrier life time also limit the performance of hematite [17]. Thus, it is highly desirable to develop new visible-light driven materials that have proper band potential positions and could simultaneously harvest a wide range of visible light. Bi4V2O11 has emerged as a visible light responsive photocatalyst and demonstrated ability for dye or organic waste degradation [18,19,20,21,22,23,24]. The visible light driven oxygen evolution performance was investigated by Jiang et al. [18] Santos et al. employed a low temperature phase as a hole inversion layer at the BiVO4/Bi4V2O11 interface [25,26]. Photocatalytic CO2 reduction was also investigated utilizing Bi-quantum-dot-decorated Bi4V2O11 hollow nanocakes [27]. Although electronic structure of orthorhombic Bi4V2O11 has been investigated by density functional theory (DFT) to predict its possibility of photoelectrochemical water splitting [28], the material has not been thoroughly investigated as a photoelectrode for PEC water splitting. Furthermore, the doping effect has not been explored to enhance the PEC performance. Doping is a proven method for increasing carrier density, which consequently enhances the charge separation of the photocatalyst [29]. The efficacy of Mo doping in enhancing the BiVO4 photocatalyst has been demonstrated [30]. This strategy, which has shown promise in BiVO4, may similarly benefit Bi4V2O11. In the present study, the properties of Bi4V2O11 photoanodes were investigated. Bi4V2O11 powders, with and without Mo-doping, were prepared by solid state reaction. The photocurrent density and stability under PEC test condition have been investigated. The Mo-doped Bi4V2O11 photoanode achieves a maximum photocurrent density of 0.3 mA/cm2 after a post deposition treatment to improve the connectivity of the drop-cast particles in the film. Bi4V2O11 displays a promising potential as a photoanode material for non-biased water splitting.

2. Experiment

2.1. Preparation of Pure and Mo Doped Bi4V2O11 Powder Photocatalyst Pristine Bi4V2O11 particles were prepared by a solid-state reaction. Stoichiometric amount of Bi2O3, NH4VO3 were mixed and ground in an agate mortar with addition of a small amount of ethanol. The mixture was then sintered at 600 °C for 3 h and 800 °C for 10 h. The obtained red powder has some impurity phase which was ground again in an agate mortar, then sintered at 800 °C for 10 h for the second time to obtain a pure phase of Bi4V2O11. Pure Bi4V2O11 particles were obtained after grinding. Mo doped Bi4V2O11 particles were synthesized by the same method, except adding suitable amount of MoO3. The doping concentration is 3% according to the optimized Mo doping level in BiVO4 reported by Luo et al. [13]. 2.2. Fabrication of Particulate Films Photoanode films were fabricated by drop casting method followed by necking treatment according to a method reported before [31]. Powder suspension was obtained by dispersing 20 mg powder into 500 μL isopropanol with sonication. The film was fabricated by dropping 50 μL suspension onto an FTO glass substrate (2 cm × 1 cm). After drying, another drop-casting process was conducted to obtain the film. The necking procedure was conducted by dropping 20 μL of ammonium metatungstate methanol solution (10 mM) on the dried film. This procedure was repeated for six times to improve the connectivity among the synthesized particles in the film. After that, half of the film was scrubbed to expose conductive substrate, then followed by heating at 500 °C for 1 h in air. 2.3. Pt Loading on the Powder Surface 1 wt% Pt was loaded as the co-catalyst by photoreduction using H2PtCl6. Typically, 0.1 g of photocatalyst, 100 mL of distilled water, 20 mL of CH3OH, and 1% amount of H2PtCl6·6H2O were placed in a glass vessel. The reactant solution was irradiated with a 300-W Xe lamp for 8 h with constant stirring. Then the Pt-loaded photocatalyst powder was filtered and washed thoroughly with deionized water and alcohol, subsequently dried at 60 °C for 12 h. 2.4. Water Splitting Test The photocatalytic H2 evolution reaction of powder samples was carried out in a Pyrex reaction vessel which was equipped into a closed gas circulation and evacuation system. The reaction was conducted in an aqueous solution containing 0.1 g of the 1 wt% Pt loaded Bi4V2O11 catalyst. Methanol (20% v/v, 100 mL) was used as the sacrificial reagent. The reaction was performed under irradiation of a 300-W Xe lamp with the light intensity of 100 mW/cm2. The evolved gas was then extracted and analysed by gas chromatography. Photoelectrochemical measurements were conducted utilizing a three-electrode setup comprising a PCI4/300™ potentiostat controlled by PHE200™ software from Gamry Electronic Instruments, Inc. The configuration included a Bi4V2O11 working electrode, a Pt foil serving as the counter electrode, and a reference electrode consisting of Ag/AgCl. A solar simulator (HAL-320, Asahi Spectra Co., Ltd., Tokyo, Japan) with an intensity of 100 mW·cm−2, calibrated using a solar reference cell, was employed as the light source for the photoelectrochemical measurements. Photocurrent density testing employs a mask with an open area of 0.28 cm2 positioned at the center. This configuration ensures greater uniformity in the film at the center to evaluate the real performance of the photoelectrode. Photocurrents associated with water oxidation were recorded in a 0.5 M Na2SO4 aqueous solution at a scan rate of 30 mV·s−1. Additionally, photocurrents were measured under hole scavenger conditions to achieve approximately 100% charge transfer efficiency, in a solution comprised of 0.5 M Na2SO4 mixed with 0.1 M Na2SO3. 2.5. Characterization The morphologies were examined via field emission scanning electron microscopy (FESEM, JEOL JSM-7600F, JEOL, Ltd., Tokyo, Japan). Crystallinity was determined through X-ray diffraction (XRD) analysis using a Shimadzu 6000 X-ray diffractometer equipped with Cu Kα radiation (λ = 0.154 nm), employing a 2θ scan mode with a fixed incidence angle of 1°. Light absorption characteristics were assessed by measuring reflectance and transmittance using a UV-visible spectrophotometer equipped with an integrating sphere (Lambda 750S, Perkin-Elmer, Shelton, CT, USA).

3. Results and Discussion

DFT calculation shows that α-Bi4V2O11 possesses a monoclinic structure with a bandgap of 2.03 eV (Figure 1a). The valence band maximum is mainly composed of the O 2p orbital, and the conduction band minimum is composed of the O 2p and V 3d states. Figure 1b shows the band structure of α-Bi4V2O11, which indicates that α-Bi4V2O11 is an indirect bandgap semiconductor.
Figure 1. (a) Calculated local partial density of states and (b) band structure of α-Bi4V2O11. Red, grey, and purple spheres in (a) represent O, V, and Bi atoms, respectively.
X-ray diffraction spectra of Bi4V2O11 powder prepared by solid state reaction indicate they belong to α-Bi4V2O11 (Powder Diffraction File No. 82-1481) without any detectable impurity (Figure 2a). Scanning electron microscopy (SEM) images of the Bi4V2O11 samples also suggest well crystallized particles with smooth surface (Figure S1). The particle size of the Bi4V2O11 samples is about several microns. Figure 2b shows that the as-synthesized red-coloured Bi4V2O11 powders exhibit visible light absorption with optical absorption edges of about 600 nm. The Tauc plot Bi4V2O11 (Figure 2c) is shown in Figure 1b. Using indirect bandgap model, calculation based on the Tauc plot suggests that Bi4V2O11 possesses a bandgap of about 1.87 eV, which is consistent with the calculated results. Such a small bandgap corresponds to a theoretical solar to hydrogen efficiency of 21.6% and a maximum solar photocurrent of ~17.6 mA/cm2 with a great potential for water splitting (Figure 2d). Bi4V2O11 film was fabricated by drop casting according to the reported method [31]. The film thickness was estimated to be a few microns according to the particle size (Figure S1). Mott-Schottky plot (Figure 3a) indicates the flat-band potential (VF) of Bi4V2O11 is −0.31 V vs. RHE. The flat-band potential represents the position of the Fermi level of the semiconductor and is generally considered to be located just under the conduction band of n-type semiconductors [32]. Hence, flat-band potential can be seen as the conduction band minimum level (CBM). Combining the bandgap value of 1.87 eV, we obtain the band position diagram of Bi4V2O11 as shown in Figure 3b. The energy levels of CBM and valence band maximum (VBM) straddle the redox potentials of H+/H2 and O2/H2O, which makes the material suitable for water splitting. Figure 3c shows the photocatalytic H2 evolution from a 20% (v/v) methanol aqueous solution (100 mL) over 1wt% Pt-loaded Bi4V2O11 powder sample (100 mg) under irradiation. Under irradiation, H2 was produced onto the surface of Pt-loaded Bi4V2O11, indicating the activity of Bi4V2O11 catalyst for H2 evolution. This confirms that CBM of Bi4V2O11 is negative than the water reduction potential.
Figure 2. (a) X-ray diffraction of as synthesized Bi4V2O11 powder by solid state reaction; (b) UV–vis light absorption spectrum of Bi4V2O11 powder obtained from powder reflectance. Inset is the digital image of the synthesized powder; (c) Tauc plots of Bi4V2O11 powder to estimate its bandgap; (d) Theoretical solar photocurrent density of Bi4V2O11 under AM 1.5G irradiation (100 mW/cm2).
Figure 3. (a) Mott Schottky plot of Bi4V2O11 film measured in 0.5 M Na2SO4 electrolyte under dark condition; (b) Diagram of the band position of Bi4V2O11 compared with water redox potential; (c) Photocatalytic H2 evolution of Bi4V2O11 powder.
The PEC performance of pristine Bi4V2O11 film was shown in Figure S2, which is very poor. Mo doping was employed to enhance the performance. Figure 4a shows SEM images of the Mo-doped Bi4V2O11 film, which reveals that the particle size is about several microns. The crystal structure remains the same after Mo doping (Figure S3). Light absorption was almost the same with pure Bi4V2O11 (Figure S4). Thus, Mo doping does not significantly change physical properties of Bi4V2O11. Figure 4b shows the photocurrents of Mo-doped Bi4V2O11 photoelectrodes, which is much higher than pure one without doping (Figure S2). This performance improvement is mainly benefited from the higher carrier density by Mo doping (Figure S5). It has been reported the particle connection was very important for the powder-fabricated thin films, and necking was necessary to improve the carrier transport between particles [9,10,31,33], since necking treatment can form effective contacts between the photocatalyst particles which facilitates the electron transport between the particles [34]. The one without necking treatment shows poor performance. Tungsten and titanium necking treatment were employed here. The photocurrent increases greatly with W necking compared with the one without, which reaches a maximum photocurrent of about 0.3 mA/cm2. However, it was found that the Ti necking treatment shows even worse performance than the one without necking. To exclude WO3 contribution to the photocurrent, the same amount of W necking agent was dropped on FTO substrate and followed by the same heating treatment as that of Mo-doped Bi4V2O11 photoelectrode. The obtained WO3 photoelectrode shows very weak photocurrent (Figure S6) with the maximum photocurrent less than 0.05 mA/cm2. We also optimized the necking treatment temperature (Figure S7) and found that 500 ºC shows the best performance. Different amount of catalysts was employed using powder suspension dropped on the FTO substrates (Figure S8), and the 2 mg/cm2 sample shows the best performance. Figure 4c shows the wavelength dependence of the incident photon to current efficiency (IPCE) of Mo-doped Bi4V2O11. The IPCE was calculated based on the measured constant potential I-t curves to avoid the transient photocurrent under monochromatic light irradiation. The IPCEs are consistent with I-V curves measured under simulated AM1.5G light, supporting the validity of the measurements. The stability of Mo-doped Bi4V2O11 photoanode (active area about 2 cm2) is tested under AM 1.5 G simulated sunlight condition (100 mW/cm−2) at 1.23 V vs. RHE. After 2 h irradiation the photocurrent density drops to about 0.24 mA, which is around 80% of the initial value (around 0.3 mA). This is a reasonably good performance in photo stability without any surface protection. The primary reason for the current drop lies in the stability of the necking agent, since WO3 is not stable in a neutral electrolyte solution. Some methods can be used to improve the stability such as coating a thin layer of protection oxides or changing the necking agent.
Figure 4. (a) SEM image of Mo-doped Bi4V2O11 film prepared by drop-casting method; (b) Photocurrents (in chop mode with light on-off repeatedly) of Mo-doped Bi4V2O11 photoelectrodes without the necking treatment and ones with W and Ti necking. Light on and off states were indicated; (c) IPCE of Mo-doped Bi4V2O11 film at 1.23 V vs. RHE; (d) Stability test of Mo-doped Bi4V2O11 film.
Discussion on the Limitation Factors of the Reported Performance The water oxidation photocurrent JH2O can be expressed by [35]:
JH20 = J0 × ηabs × ηsep × ηtran = Jabs × ηsep × ηtran
where J0 is the theoretical solar photocurrent density which assumes that all the solar energy corresponding to the band edge can be fully converted to fuel energy [36]. The charge transfer efficiency can be approximated to be 100% (ηinj = 1) when an effective hole scavenger Na2SO3 with fast oxidation kinetics is utilized [35]. The Na2SO3 oxidation photocurrent can be determined by [36]:
JNa2O3 = J0 × ηabs × ηsep = Jabs × ηsep
where JNa₂SO₃ is the oxidation photocurrent using Na2SO3. Combining Equations (1) and (2), we obtain the charge separation efficiency ηsep = JNa₂SO₃/Jabs and the charge transfer efficiency ηtran = JH₂O/JNa₂SO₃. Figure 5a shows the light absorption efficiency of the as-fabricated film. The calculated Jabs by integrating the light absorption efficiency over the solar spectrum (AM 1.5 G) is estimated to be 7.6 mA/cm2. Figure 5b and 5c show the calculated charge separation efficiency, ηsep, and the charge transfer efficiency, ηtran, which are about 3% and 60% at 1.23 V vs. RHE, respectively. Thus, poor charge separation is the limitation factor, which results in the low photocurrent performance. The poor charge separation is mainly limited by the fabrication method through two aspects: (1) big particle size and the corresponding thick film have unnecessarily extended distance that the carriers have to travel, and thus increased the chance of recombination; (2) weak adhered boundaries between particles may have resulted in severe recombination. To improve the performance, more effort is needed to reduce the particle size by either exploring low temperature synthesis method for finer particles or making use vapor deposition method to prepare the film directly.
Figure 5. (a) Light absorption efficiency (b) charge separation efficiency and (c) charge transfer efficiency of Mo doped Bi4V2O11 film.

4. Conclusions

Bi4V2O11 possesses a red color with a bandgap of ~1.9 eV, corresponding to a highly promising solar to hydrogen efficiency. The CBM is more negative than water reduction potential of H+/H2 and VBM is more positive than the water oxidation potential of O2/H2O. This has made the material attractive for non-biased water splitting. Pristine Bi4V2O11 films show quite poor performance, and Mo doping can greatly enhance the photocurrent by enhancing the charge carrier density. W necking was necessary to lead to a better performance in our case of films consisting of relatively large particles. The maximum photocurrent of Mo-doped Bi4V2O11 photoanode reaches a value of about 0.3 mA/cm2 with W necking treatment. The Mo-doped Bi4V2O11 photoanode maintains 80% photocurrent after 2 h test, displaying reasonably good stability among known visible light active photocatalysts. The reason for the limited PEC performance has been analysed with potential strategies proposed to overcome the limitations for this promising photoanode material.

Supplementary Materials

The following supporting information can be found at: https://www.sciepublish.com/article/pii/167, DFT calculation method; Figure S1. SEM images of as synthesized Bi4V2O11 particles; Figure S2. Photocurrent of pristine Bi4V2O11 photoelectrode with W necking treatment; Figure S3. X-ray diffraction of Mo doped Bi4V2O11 powder by solid state reaction; Figure S4. UV-vis light absorption spectrum of Mo doped Bi4V2O11 powder; Figure S5. Mott Schottky plot of Mo doped Bi4V2O11 film measured in 0.5 M Na2SO4 electrolyte under dark condition; Figure S6. Photocurrent of WO3 obtained by dropping the same amount as necking agent on FTO substrate and heated at 500 °C for 1h; Figure S7. Photocurrent of W necking samples treated with different temperature; Figure S8. Photocurrent of samples with different amount of powder suspension dropped on the FTO substrates.

Author Contributions

Conceptualization, J.H., J.F., and Z.C.; Methodology, X.Z. and J.F.; Validation, X.Z., J.H., J.F.; Formal Analysis, X.Z. and Y.Y.; Investigation, X.Z., N.Z., T.F. and J.F.; Resources, Z.C. and J.F.; Data Curation, J.H. and J.F.; Writing—Original Draft Preparation, Z.X. and Y.Y.; Writing—Review & Editing, Z.C.; Supervision, J.H., J.F. and Z.C.; Project Administration, Z.C.; Funding Acquisition, Z.C.

Ethics Statement

Not applicable.

Informed Consent Statement

Not applicable.

Funding

This research was funded by the National Research Foundation, Singapore for funding through a CREATE Program.

Declaration of Competing Interest

The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.

References

1.
Wang Z, Wu Q, Wang J, Yu Y. Surface modification by ligand growth strategy for dense copper bismuth film as photocathode to enhance hydrogen production activity. Front. Energy 2023, doi:10.1007/s11708-023-0893-5.
2.
Jun SE, Kim Y-H, Kim J, Cheon WS, Choi S, Yang J, et al. Atomically dispersed iridium catalysts on silicon photoanode for efficient photoelectrochemical water splitting. Nat. Commun. 2023, 14, 609. [Google Scholar]
3.
Wang X, Gong J, Dong Y, An S, Zhang X, Tian J. Energy band engineering of hydroxyethyl group grafted on the edge of 3D g-C3N4 nanotubes for enhanced photocatalytic H2 production. Mater. Today Phys. 2022, 27, 100806. [Google Scholar]
4.
Díez-García MI, Gómez R. Progress in ternary metal oxides as photocathodes for water splitting cells: optimization strategies.  Sol. RRL 2022, 6, 2100871. [Google Scholar]
5.
Hitoki G, Ishikawa A, Takata T, Kondo JN, Hara M, Domen K. Ta3N5 as a novel visible light-driven photocatalyst (λ < 600 nm). Chem. Lett. 2002, 31, 736–737. [Google Scholar]
6.
Kawase Y, Higashi T, Obata K, Sasaki Y, Katayama M, Domen K, et al. Interfacial design of a Ta3N5 thin-film photoanode for highly stable oxygen evolution over a wide pH range. ACS Sustain. Chem. Eng. 2022, 10, 14705–14714. [Google Scholar]
7.
Hara M, Hitoki G, Takata T, Kondo JN, Kobayashi H, Domen K. TaON and Ta3N5 as new visible light driven photocatalysts. Catal. Today 2003, 78, 555–560. [Google Scholar]
8.
Minegishi T, Nishimura N, Kubota J, Domen K. Photoelectrochemical properties of LaTiO2N electrodes prepared by particle transfer for sunlight-driven water splitting. Chem. Sci. 2013, 4, 1120–1124. [Google Scholar]
9.
Feng J, Luo W, Fang T, Lv H, Wang Z, Gao J, et al. Highly photo-responsive LaTiO2N photoanodes by improvement of charge carrier transport among film particles. Adv. Funct. Mater. 2014, 24, 3535–3542. [Google Scholar]
10.
Zhong Y, Li Z, Zhao X, Fang T, Huang H, Qian Q, et al. Enhanced water-splitting performance of perovskite SrTaO2N photoanode film through ameliorating interparticle charge transport. Adv. Funct. Mater. 2016, 26, 7156–7163. [Google Scholar]
11.
Kim Y-I, Si W, Woodward PM, Sutter E, Park S, Vogt T. Epitaxial thin-film deposition and dielectric properties of the perovskite oxynitride BaTaO2N. Chem. Mater. 2007, 19, 618–623. [Google Scholar]
12.
Li Z, Luo W, Zhang M, Feng J, Zou Z. Photoelectrochemical cells for solar hydrogen production: Current state of promising photoelectrodes, methods to improve their properties, and outlook. Energy Environ. Sci. 2013, 6, 347–370. [Google Scholar]
13.
Luo W, Yang Z, Li Z, Zhang J, Liu J, Zhao Z, et al. Solar hydrogen generation from seawater with a modified BiVO4 photoanode. Energy Environ. Sci. 2011, 4, 4046–4051. [Google Scholar]
14.
Alexander BD, Kulesza PJ, Rutkowska I, Solarska R, Augustynski J. Metal oxide photoanodes for solar hydrogen production. J. Mater. Chem. 2008, 18, 2298–2303. [Google Scholar]
15.
Fujishima A, Honda K. Electrochemical photolysis of water at a semiconductor electrode. Nature 1972, 238, 37–38. [Google Scholar]
16.
Cho IS, Han HS, Logar M, Park J, Zheng X. Enhancing low-bias performance of hematite photoanodes for solar water splitting by simultaneous reduction of bulk, interface, and surface recombination pathways. Adv. Energy Mater. 2016, 6, 1501840. [Google Scholar]
17.
Piekner Y, Ellis DS, Grave DA, Tsyganok A, Rothschild A. Wasted photons: Photogeneration yield and charge carrier collection efficiency of hematite photoanodes for photoelectrochemical water splitting. Energy Environ. Sci. 2021, 14, 4584–4598. [Google Scholar]
18.
Chen X, Liu J, Wang H, Ding Y, Sun Y, Yan H. One-step approach to novel Bi4V2O11 hierarchical hollow microspheres with high visible-light-driven photocatalytic activities. J. Mater. Chem. A 2013, 1, 877–883. [Google Scholar]
19.
Kumar S, Sahare PD. Photocatalytic activity of bismuth vanadate for the degradation of organic compounds. Nano 2013, 08, 1350007. [Google Scholar]
20.
Lv C, Chen G, Sun J, Zhou Y. Construction of α–β phase junction on Bi4V2O11 via electrospinning retardation effect and its promoted photocatalytic performance. Inorg. Chem. 2016, 55, 4782–4789. [Google Scholar]
21.
Ri C-N, Song-Gol K, Ju-Yong J, Pak S-N, Ri S-C, Ri J-H. Construction of the Bi2WO6/Bi4V2O11 heterojunction for highly efficient visible-light-driven photocatalytic reduction of Cr(vi). New J. Chem. 2018, 42, 647–653. [Google Scholar]
22.
Wen X-J, Qian L, Lv X-X, Sun J, Guo J, Fei Z-H, et al. Photocatalytic degradation of sulfamethazine using a direct Z-Scheme AgI/Bi4V2O11 photocatalyst: Mineralization activity, degradation pathways and promoted charge separation mechanism. J. Hazard. Mater. 2020, 385, 121508. [Google Scholar]
23.
Li J, Lu P, Deng W, Zeng Z, Lin L, Zhao G. Facile synthesis of sheet-like BiVO4/Bi4V2O11 composite for enhanced photocatalytic properties. Mater. Chem. Phys. 2020, 254, 123489. [Google Scholar]
24.
Lin Y, Cai H, Chen H, Luo H. One-pot synthesis of Bi4V2O11/BiVO4 heterostructure with enhanced photocatalytic activity for dye degradation. Appl. Surf. Sci. 2021, 544, 148921. [Google Scholar]
25.
dos Santos WS, Rodriguez M, Afonso AS, Mesquita JP, Nascimento LL, Patrocínio AOT, et al. A hole inversion layer at the BiVO4/Bi4V2O11 interface produces a high tunable photovoltage for water splitting. Sci. Rep. 2016, 6, 31406. [Google Scholar]
26.
dos Santos WS, Rodriguez M, Khoury JMO, Nascimento LA, Ribeiro RJP, Mesquita JP, et al. Bismuth vanadate photoelectrodes with high photovoltage as photoanode and photocathode in photoelectrochemical cells for water splitting. ChemSusChem 2018, 11, 589–597. [Google Scholar]
27.
Zhao X, Duan Z, Chen L. Bi-quantum-dot-decorated Bi4V2O11 hollow nanocakes: Synthesis, characterization, and application as photocatalysts for CO2 reduction. Ind. Eng. Chem. Res. 2019, 58, 10402–10409. [Google Scholar]
28.
Liu J, Ozawa K, Uezono N, Pawar SA, Suzuki S, Traoré A, et al. Electronic structure of orthorhombic Bi4V2O11 and role of gap states in photoelectrochemical water splitting. J. Chem. Phys. C 2023, 127, 11195–11203. [Google Scholar]
29.
Wang Y, Liang S, Zuo C, Fang H, Dong G, Sheng X, et al. Construction of a heterojunction with fast charge transport channels for photocatalytic hydrogen evolution via a synergistic strategy of Co-doping and crystal plane modulation. Nanoscale 2023, 15, 5230–5240. [Google Scholar]
30.
Wang Z, Zhang W, Song Y, Liu N, Chen L, An N, et al. Unraveling the site-selective doping mechanism in single-crystalline BiVO4 thin films for photoelectrochemical water splitting. J. Chem. Phys. C 2023, 127, 5775–5782. [Google Scholar]
31.
Feng J, Zhao X, Ma SSK, Wang D, Chen Z, Huang Y. Fast and simple construction of efficient solar-water-splitting electrodes with micrometer-sized light-absorbing precursor particles. Adv. Mater. Technol. 2016, 1, 1600119. [Google Scholar]
32.
Sayama K, Nomura A, Arai T, Sugita T, Abe R, Yanagida M, et al. Photoelectrochemical decomposition of water into H2 and O2 on porous BiVO4 thin-film electrodes under visible light and significant effect of Ag ion treatment. J. Phys. Chem. B 2006, 110, 11352–11360. [Google Scholar]
33.
Zhang L, Song Y, Feng J, Fang T, Zhong Y, Li Z, et al. Photoelectrochemical water oxidation of LaTaON2 under visible-light irradiation. Int. J. Hydrogen Energy 2014, 39, 7697–7704. [Google Scholar]
34.
Higashi M, Domen K, Abe R. Fabrication of efficient TaON and Ta3N5 photoanodes for water splitting under visible light irradiation. Energy Environ. Sci. 2011, 4, 4138–4147. [Google Scholar]
35.
Dotan H, Sivula K, Grätzel M, Rothschild A, Warren SC. Probing the photoelectrochemical properties of hematite (α-Fe2O3) electrodes using hydrogen peroxide as a hole scavenger. Energy Environ. Sci. 2011, 4, 958–964. [Google Scholar]
36.
Zhao X, Luo W, Feng J, Li M, Li Z, Yu T, et al. Quantitative analysis and visualized evidence for high charge separation efficiency in a solid-liquid bulk heterojunction. Adv. Energy Mater. 2014, 4, 1301785. [Google Scholar]
TOP