Next Article in Journal
Simultaneous Effect of Diameter and Concentration of Multi-Walled Carbon Nanotubes on Mechanical and Electrical Properties of Cement Mortars: With and without Biosilica
Previous Article in Journal
Thermoplastic Polyurethane Derived from CO2 for the Cathode Binder in Li-CO2 Battery
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Oxygen-Vacancy-Induced Enhancement of BiVO4 Bifunctional Photoelectrochemical Activity for Overall Water Splitting

1
School of Physics and Technology, Nantong University, Nantong 226019, China
2
Research Center for Quantum Physics and Materials, Nantong University, Nantong 226019, China
3
College of Physics Science and Technology, Yangzhou University, Yangzhou 225002, China
*
Authors to whom correspondence should be addressed.
Nanomaterials 2024, 14(15), 1270; https://doi.org/10.3390/nano14151270
Submission received: 17 June 2024 / Revised: 24 July 2024 / Accepted: 26 July 2024 / Published: 29 July 2024

Abstract

:
Hydrogen generation via photoelectrochemical (PEC) overall water splitting is an attractive means of renewable energy production so developing and designing the cost-effective and high-activity bifunctional PEC catalysts both for the hydrogen evolution reaction (HER) and the oxygen evolution reaction (OER) has been focused on. Based on first-principles calculations, we propose a feasible strategy to enhance either HER or OER performance in the monoclinic exposed BiVO4 (110) facet by the introduction of oxygen vacancies (Ovacs). Our results show that oxygen vacancies induce charge rearrangements, which enhances charge transfer between active sites and adatoms. Furthermore, the incorporation of oxygen vacancies reduces the work function of the system, which makes charge transfer from the inner to the surface more easily; thus, the charges possess stronger redox capacity. As a result, the Ovac reduces both the hydrogen adsorption-free energy (ΔGH*) for the HER and the overpotential for the OER, facilitating the PEC activity of overall water splitting. The findings provide not only a method to develop bifunctional PEC catalysts based on BiVO4 but also insight into the mechanism of enhanced catalytic performance.

1. Introduction

Driven by growing concerns about environmental pollution and the energy crisis, developing renewable “green” energy has become one of the most concerning scientific topics [1,2]. Photoelectrochemical (PEC) water splitting is one of the most feasible approaches to generating hydrogen and oxygen using solar energy. The photogenerated charge carriers involve a water redox reaction at the catalyst surface [3,4,5]. There are two electrolysis steps in the water-splitting devices, the hydrogen evolution reaction (HER) at the cathode and the oxygen evolution reaction (OER) at the anode. The HER includes two electron-transfer steps [6] while the OER is more sluggish with an energetically uphill four-electron transfer [7,8]. During the water-splitting process, ionized hydrogen ions are equivalent to the hydroxide ions. Therefore, hydrogen and oxygen can be produced through PEC overall water splitting at the same time without adding a sacrificial agent [9,10]. However, it is difficult to search for spontaneous and efficient catalysts for overall water splitting because they should not only satisfy the stringent requirements for PEC water splitting but also show efficient HER and OER behaviors [11,12,13,14].
Several bifunctional catalysts have been designed and, though some of them can release holes and excite electrons to oxidize water and reduce photons, their catalytic activity is poor [15,16,17]. Additionally, some exhibit good behavior in a single reaction while improving the performance in the other reaction [18,19,20]. Developing promising and efficient semiconductor bifunctional photocatalysts remains a great challenge. Recently, monoclinic clinobisvanite bismuth scheelite (ms-BiVO4) has received extensive concern due to the ideal gap (2.4 eV) for visible-light absorption, suitable valence band edge position (~2.8 V vs. RHE), and superior hole mobility [21,22]. As we know, BiVO4 single crystals with a different fraction of exposed facets have been synthesized, exhibiting different PEC activity [23,24,25]. In this case, the (110) facet is highly exposed with good stability and is easy to separate and accumulate photogenerated charges. Nevertheless, the PEC performance of undecorated BiVO4 is mediocre due to its shortcomings in fast surface electron-hole recombination and sluggish water redox behaviors [26,27]. Various improvement strategies, such as combining heterostructure catalysts, doping with impurities, and so on [28,29,30,31,32], have been adopted to overcome these shortcomings. In this case, oxygen vacancy modification had a beneficial influence on water oxidation in BiVO4 owing to creating more active sites and favoring charge separation and transfer [33,34,35,36]. There are many ways to induce and increase oxygen vacancies in BiVO4. For example, Xu et al. applied an ionized argon plasma technology on three-dimensional nanoporous BiVO4 to controllably generate surface oxygen vacancies [23]. Mayur et al. demonstrated a rapid Fenton-like reaction method for fabricating an ultrathin amorphous Ni:FeOOH nanolayer with an in-situ-induced Ovac on the BiVO4 photoanode [24].
In previous work, BiVO4 (110) facet modified by oxygen vacancies has been investigated to meet the requirements of PEC water splitting [37], which have been verified by experiments [23,24,25]. Under irradiation, photogenerated electrons and holes can provide driving forces for the HER and OER. Though some works have investigated the effect of the Ovac on the BiVO4, they mainly discussed the stability and electronic structures. In this work, we focus on the effect of the Ovac on bifunctional PEC activity, simultaneously, for the HER and OER in the BiVO4 (110) facet, not only analyzing the electric structure but also describing the thermodynamic process for the HER and OER. The results show that oxygen vacancy not only decreases the hydrogen adsorption free energy (ΔGH*) of the HER but also reduces the overpotential of the OER; this is mainly because the Ovac induces charge rearrangement to enhance charge transfer from the active site to adsorbate atoms. Interestingly, during the water-splitting process, under the interference of hydrogen and oxygen, more photogenerated charges can be excited. Additionally, the Ovac reduces the work function and favors the carrier’s transition from the inner to the surface. As a result, the introduction of the Ovac can effectively improve bifunctional PEC activity in the BiVO4 (110) facet.

2. Computational Model and Methods

To investigate the HER and OER performance of the BiVO4 (110) facet, the Vienna ab initio simulation package (VASP) was employed by the density functional theory (DFT). The generalized gradient approximation (GGA) was used with Perdew–Burke–Ernzerhof (PBE) [38,39,40,41], the energy cutoff was 400 eV, the convergent criterion was an energy change less than 1.0 × 10−5 eV, and the optimization was completed when the maximum forces were 0.01 eV·Å−1. The k-points were set to 5 × 5 × 1 for geometric optimization and 7 × 7 × 1 for electronic structure calculations. We compared the GGA-calculated lattice parameters to HSE methods. The GGA-calculated lattice parameters a = 5.04 Å, b = 5.27 Å, and c = 11.89 Å were well matched with the experimental values (a = 5.10 Å, b = 5.17 Å, and c = 11.69 Å) [42,43]. Compared to the HSE06-calculated band gap of 2.80 eV [44], the band gap obtained by the GGA method of 2.25 eV was closer to the experimental value of 2.45 eV [43], proving the reliability of the GGA method. The bilayer (110) facet was cleaved from the optimized ms-BiVO4 bulk and a 12 Å vacuum slab was added along the z-direction to minimize the potential artificial interaction between the adjacent images, as shown in Figure 1a. To investigate the effect of the Ovac on the active atom, we built the Ovac neighboring to the active Bi atom, as shown in Figure 1b.

3. Results and Discussion

3.1. The Effect of the Ovac on the Electronic Structure of the BiVO4 (110) Facet

In our previous work, we discussed the effect of the Ovac on BiVO4 in detail. The formation energy can be defined as Eform = EOvacEsurf + 1/2 EO2, where EOvac, Esurf, and EO2 are the total energies of BiVO4 (110) facets with and without the Ovac and molecular O2, respectively. The formation energy is 3.86 eV and the calculated result is comparable to the previous report, indicating oxygen vacancies can be easily formed in the BiVO4 (110) facet [37]. Additionally, the Ovac induces some changes in the structure. It is these changes that result in the change in electrical structure. As shown in Figure 1, BiVO4 (110) facets without and with the Ovac are all semiconductors with band gaps of 2.28 eV and 2.33 eV. Similarly, the VBM mainly consists of V 3d and O 2p and the CBM is primarily composed of Bi 6p, O 2p, and V 3d. Differently, owing to the presence of the Ovac, electron density redistributes and localized states appear in the band gap of BiVO4, whose components are V 3d and O 2p, promoting charge carrier concentration and mobility [37]. The work function decreases from 6.04 eV to 5.87 eV, the carrier’s transition from the solid to the vacuum region becomes easy, and, therefore, more photogenerated carriers will participate in catalytic reactions. More importantly, owing to the Ovac-induced internal electric field, the CBM upshifts and straddles the water reduction potential. As a result, the BiVO4 (110) facet satisfies the requirements of PEC water splitting by introducing the Ovac; the detailed analysis of electronic structure, optical adsorption, and band edge alignment can be found in Ref. [37]. Photogenerated electrons and holes can provide the charges for further HERs and OERs.

3.2. The Effect of the Ovac on the HER of the BiVO4 (110) Facet

As we know, the acidity and alkalinity of the solution will affect catalytic performance. Compared to acid HER behavior, the alkaline HER is more complex. Here, we focus on discussing the effect of the Ovac on active sites; thus, PEC water splitting in acidic conditions is considered. During the HER process, hydrogen adsorption on the BiVO4 surface takes place; the adsorption-free energy (ΔGH*) is an important descriptor for evaluating catalytic activity, which can be defined as ΔGH* = EH + ΔEZPETΔSH, where EH is hydrogen adsorption energy, ΔEZPE is the difference in zero-point energy of hydrogen vibration between the adsorbed state and the gas phase, and ΔSH is the entropy difference between the adsorbed state of the system and gas phase at the standard condition [45]. The ideal ΔGH* is nearly zero because the weak binding (ΔGH* > 0) reduces the sustainability of the reaction and the strong binding (ΔGH* < 0) makes hydrogen difficult to capture from the catalysts [46]. In the BiVO4 (110) facet, the surface O atom is the active site for the HER and surface different O atoms, respectively, act as active sites for H adsorption in Figure 1. The ΔGH* values are −0.58, −0.49, and −0.56 eV for O1, O2, and O3, respectively; the more negative values indicate stronger binding between the active site and H* but strong binding makes it difficult for hydrogen ions to separate from the surface; thus, O2 is considered a HER active site for further investigation.
As shown in Figure 2, with the assistance of the Ovac, the ΔGH* of the active O2 site decreases to −0.39 eV, indicating the Ovac weakens the binding between the H adatom and surface, and hydrogen dissociation becomes easy. This is because the appearance of the Ovac induces unsaturated bonds and unsaturated electrons, which leads to charge rearrangement; as a result, there is 7.19 e on the active site of O2 on the (110)-Ovac surface, slightly more than that on the (110) surface with 7.16 e based on Bader charge analysis. Compared to Figure 2a, Figure 2b shows that shallow energy levels appear near the VBM, mainly deriving from V and O atoms; this shallow level should be the donor level introduced by the n-doping, as shown by a peak above the Fermi level. Importantly, owing to H intervention, more states appear at intermediate levels in the H-adsorbed (110)-Ovac facet than the (110)-Ovac facet, meaning H not only participates in the HER but also promotes the generation of photogenerated charges, thus benefiting optical adsorption enhancement.

3.3. The Effect of the Ovac on the OER of the BiVO4 (110) Facet

For the OER, there are four electron-transfer steps, more complicated than the HER, and each step involves electron transfer accompanied by proton removal:
H 2 O   ( l ) +   HO +   H + +   e ,
HO O +   H + +   e ,
O +   H 2 O HOO +   H + +   e ,
HOO O 2 ( g ) +   H + +   e ,
where * represents the surface adsorption site.
As we know, the active site of BiVO4 for the OER is the surface Bi atom; four different Bi atoms of the surface are considered H2O* adsorption sites and the adsorption energies are defined as Eads = EtotEsurfE(*) (where Etot, Esurf, and E(*) are total energies of (110) facets with and without adsorbate and the energy of adsorbate), which are −0.60, −0.36, −0.46, and −0.67 eV for Bi1, Bi2, Bi3, and Bi4 atoms, respectively. All are exothermic and the binding strength between the Bi4 atom and surface is more energetically favorable; thus, Bi4 is considered the active site of the OER for further investigation. During the first reaction step, H2O* is adsorbed on top of the Bi atom; the Bi–O bond length is 2.57 Å and the H2O* adsorption energy is −0.67 eV in the (110) facet. For the (110)-Ovac facet, the bond length of Bi–O decreases to 2.44 Å and the Eads of H2O* decreases to −0.82 eV. The changes come from the Ovac-induced charge redistribution; a more unsaturated surface is more hydrophilic, displaying stronger interaction between H2O* and the surface [47]. Compared to Figure 3a, more impurity levels appear at the VBM and CBM in Figure 3b, reducing the band gap; additionally, the impurity levels also appear at the band gap for O* and HOO* adsorbed on BiVO4 (110). Additionally, the PDOS of H2O* adsorption in Figure 3 shows great changes have taken place not only on Bi atoms but also on V atoms in the (110)-Ovac facet. The charge density difference (see inset in Figure 3) further displays great charge transfer in the (110)-Ovac facet occurs not only on the surface Bi atom but also on the neighboring V atom, indicating the adjacent atoms have a great effect on the catalytic reaction owing to the existence of the Ovac. Additionally, the isolated states on the band gap move to the conduction band when the (110)-Ovac facet adsorbs H2O*, indicating that H2O is easily adsorbed on the surface owing to the existence of the Ovac. Similar to H2O* adsorption, the bond lengths of Bi–O decrease greatly to 2.109, 2.164, and 2.193 Å for HO*, O*, and HOO* adsorption in the (110)-Ovac facet while they are 2.265, 2.280, and 2.636 Å in the pure (110) facet. The shorter bond lengths indicate stronger interaction between the surface Bi atom and O atom and the charge transfer from surface Bi to the adsorbed O atom becomes more easy.
In addition, the Ovac-induced states are significantly enhanced for the HO*, O*, and HOO* adsorbed in the (110)-Ovac facet, indicating more photogenerated charges can be excited in the OER process, which can be verified by work functions in Figure 4. The work functions are 6.01, 6.63, 6.25, and 5.75 eV for H2O*, HO*, O*, and HOO* adsorption in the pure (110) facet; they reduce greatly to 4.05, 4.32, 4.77, and 4.82 eV in the (110)-Ovac facet. The reduced work functions show the charges are more easily transferred from bulk to surface in each electron step; simultaneously, the charges possess higher energies and stronger oxidation capacity, agreeing with experimental results that the Ovac favors charge separation and transfer [23].
To gain insight into the thermodynamics and direction of the reaction, we calculate the free energy diagrams of the OER process on BiVO4 (110) facets. The free energy can be defined as ΔG = ΔE + ΔZPE-TΔS, in which ΔE is the reaction energy depending on each electron-transfer step, ΔZPE denotes the change of zero point energies, and ΔS is the change in entropic contribution by employing the computed vibrational frequencies and standard tables for the reactants and products in the gas phase [48,49]. In the free energy calculation, an external bias U can be applied on each electron-transfer step. The free energies can be expressed as:
Δ G 1 = E ( HO ) E surf E H 2 O + 1 2 E H 2 + ( Δ ZPE T Δ S ) e U
Δ G 2 = E ( O ) E ( HO ) + 1 2 E H 2 + ( Δ ZPE T Δ S ) e U
Δ G 3 = E ( HOO ) E ( O ) E H 2 O + 1 2 E H 2 + ( Δ ZPE T Δ S ) e U
Δ G 4 = E surf E ( HOO ) + E O 2 + 1 2 E H 2 + ( Δ ZPE T Δ S ) e U
The total reaction H 2 O 1 2 O 2 +   H 2 occurs at the potential of 2.46 V per water molecule; the minimum free energy should split two H2O molecular with 4.92 V [50].
Figure 5a shows the free energies of the OER process in the BiVO4 (110) facet at U = 0 V, pH = 0, and T = 298 K; the first step is the adsorption of H2O* moiety on the surface Bi site and removing a proton from H2O* to create an HO* radical and the free energy runs steeply uphill to 2.66 eV. This process consumes more energy than other processes, which is a rate limitation, indicating water dissociation is difficult. The reaction continues and the HO* moiety releases another proton generating O*; the free energy is 3.86 eV. The O* is very electrophilic and immediately gains an electron by bonding with adjacent H2O* and forming HOO* with the free energy of 5.34 eV. Lastly, O2 is released from the (110) surface. The free energies of the four processes are all uphill; it is necessary to apply potential bias (or overpotential) to make each step downhill. The overpotential can be obtained by the difference between the voltage for all free-energy steps downhill and the minimum voltage required for the OER [51]. For the (110) facet, the first step to form the HO* radical is very critical to determine the limiting potential with the overpotential of 1.43 V (=2.66 − 1.23 V). Importantly, the overpotential of the (110)-Ovac facet decreases greatly with 0.62 V and the third step for generating the HOO* radical becomes the determining-potential step (see Figure 5b).
The reduction in overpotential originates from charge rearrangement, which can be found from the Bader charge on the active site Bi at each electron step in Figure 6. For the pure BiVO4 (110) facet, the active site Bi atom has the most charges at the first step with 3.22 e, displaying weak oxidation properties, which limits the reaction; thus, the first step is the rate-determining step. Meanwhile for the (110)-Ovac facet, the charges on the Bi atom overall decrease, indicating greater charge transfer from active Bi to neighboring atoms. Additionally, the Bi atom has the most charges with 3.10 e at the third step, which becomes the rate-control step. Obviously, the presence of the Ovac induces charge rearrangement and accelerates charge transfer then effectively enhances the activity of oxygen generation.
Furthermore, we investigate the effect of the Ovac position and the Ovac concentration on the OER. As shown in Figure 7a, we create the Ovac coordinated with V, comparing with the Ovac coordinated with Bi (see Figure 5b); the second step becomes the rate-control step with the overpotential of 0.71 V. Additionally, we create two and three Ovacs to increase the Ovac concentration to 6.25% and 9.38%, as shown in Figure 7b,c. The rate control steps occur at the fourth step with the overpotential of 0.32 V for a 6.25% Ovac concentration and the third step with the overpotential of 0.63 V for a 9.38% Ovac concentration. These indicate that the Ovac site and concentration have a significant influence on OER properties.

4. Conclusions

Based on the DFT calculation, we investigate the effect of the Ovac on the catalytic activity of overall water splitting in the BiVO4 (110) facet. The results have shown that the Ovac induces charge rearrangement; there appear intermediate shallow levels, favoring charge transfer from the VBM to the CBM. During the HER and OER processes, the intervention of H and O promotes the generation of photogenerated charges. Meanwhile, there occurs great charge transfer between the active site and neighboring atoms, which not only simulates the neighboring atom activity but also enhances the interaction between the active sites. Additionally, the Ovac reduces the work function, which not only benefits theoverflow of charge but also enhances charge oxidation capacity. Therefore, the (110)-Ovac facet has lower ΔGH* for the HER and the overpotential for the OER. Contributing from the Ovac, bifunctional catalytic activity can be effectively improved in the BiVO4 (110) facet. Importantly, vacancy-defect engineering is a feasible strategy to improve the PEC water-splitting activity of BiVO4.

Author Contributions

Methodology, Y.L.; Investigation, Q.Q.; Writing—original draft, H.F.; Writing—review & editing, J.P.; Funding acquisition, C.Z. All authors have read and agreed to the published version of the manuscript.

Funding

This work is supported by grants from the National Natural Science Foundation of China (No 11974354, 12074332). The authors are grateful for access to the computational resources at NTU.

Data Availability Statement

The data that support the findings of this study are available within the article.

Conflicts of Interest

The authors have no conflicts of interest to disclose.

References

  1. Cao, M.; Wang, X.; Cao, W.; Fang, X.; Wen, B.; Yuan, J. Thermally Driven Transport and Relaxation Switching Self-Powered Electromagnetic Energy Conversion. Small 2018, 14, 1800987. [Google Scholar] [CrossRef] [PubMed]
  2. Li, L.; Chang, X.; Lin, X.; Zhao, Z.J.; Gong, J. Theoretical insights into single-atom catalysts. Chem. Soc. Rev. 2020, 49, 8156–8178. [Google Scholar] [CrossRef] [PubMed]
  3. Wang, H.; Xia, Y.; Li, H.; Wang, X.; Yu, Y.; Jiao, X.; Chen, D. Highly active deficient ternary sulfide photoanode for photoelectrochemical water splitting. Nat. Commun. 2020, 11, 3078. [Google Scholar] [CrossRef] [PubMed]
  4. Zhang, E.; Liu, Z.; Kato, M. Durable and efficient photoelectrochemical water splitting using TiO2 and 3C–SiC single crystals in a tandem structure. Sol. Energy Mater. Sol. Cells 2021, 230, 111260. [Google Scholar] [CrossRef]
  5. Low, J.; Yu, J.; Jaroniec, M.; Wageh, S.; Al-Ghamdi, A.A. Heterojunction Photocatalysts. Adv. Mater. 2017, 29, 1601694. [Google Scholar] [CrossRef] [PubMed]
  6. Han, S.; Noh, S.; Yu, Y.T.; Lee, C.R.; Lee, S.K.; Kim, J.S. Highly Efficient Photoelectrochemical Water Splitting Using GaN-Nanowire Photoanode with Tungsten Sulfides. ACS Appl. Mater. Interfaces 2020, 12, 58028–58037. [Google Scholar] [CrossRef] [PubMed]
  7. Cui, X.; Ren, P.; Ma, C.; Zhao, J.; Chen, R.; Chen, S.; Rajan, N.P.; Li, H.; Yu, L.; Tian, Z.; et al. Robust Interface Ru Centers for High-Performance Acidic Oxygen Evolution. Adv. Mater. 2020, 32, 1908126. [Google Scholar] [CrossRef] [PubMed]
  8. Sun, Y.; Xue, Z.; Liu, Q.; Jia, Y.; Li, Y.; Liu, K.; Lin, Y.; Liu, M.; Li, G.; Su, C.Y. Modulating electronic structure of metal-organic frameworks by introducing atomically dispersed Ru for efficient hydrogen evolution. Nat. Commun. 2021, 12, 1369. [Google Scholar] [CrossRef] [PubMed]
  9. Hu, E.; Feng, Y.; Nai, J.; Zhao, D.; Hu, Y.; Lou, X.W. Construction of hierarchical Ni–Co–P hollow nanobricks with oriented nanosheets for efficient overall water splitting. Energy Environ. Sci. 2018, 11, 872–880. [Google Scholar] [CrossRef]
  10. Xu, Z.; Ying, Y.; Zhang, G.; Li, K.; Liu, Y.; Fu, N.; Guo, X.; Yu, F.; Huang, H. Engineering nife layered double hydroxide by valence control and intermediate stabilization toward the oxygen evolution reaction. J. Mater. Chem. A 2020, 8, 26130–26138. [Google Scholar] [CrossRef]
  11. Kalanur, S.S.; Singh, R.; Seo, H. Enhanced solar water splitting of an ideally doped and work function tuned {002} oriented one-dimensional WO3 with nanoscale surface charge mapping insights. Appl. Catal. B Environ. 2021, 295, 12029. [Google Scholar] [CrossRef]
  12. Xie, Z.; Chen, D.; Zhai, J.; Huang, Y.; Ji, H. Charge separation via synergy of homojunction and electrocatalyst in BiVO4 for photoelectrochemical water splitting. Appl. Catal. B Environ. 2023, 334, 122865. [Google Scholar] [CrossRef]
  13. Montoya, J.H.; Seitz, L.C.; Chakthranont, P.; Vojvodic, A.; Jaramillo, T.F.; Norskov, J.K. Materials for solar fuels and chemicals. Nat. Mater. 2016, 16, 70–81. [Google Scholar] [CrossRef] [PubMed]
  14. Kalanur, S.S.; Lee, Y.J.; Seo, H. A versatile synthesis strategy and band insights of monoclinic clinobisvanite BIVO4 thin films for enhanced photoelectrochemical water splitting activity. Appl. Surf. Sci. 2021, 562, 150078. [Google Scholar] [CrossRef]
  15. Hu, P.; Jia, Z.; Che, H.; Zhou, W.; Liu, N.; Li, F.; Wang, J. Engineering hybrid CoMoS4/Ni3S2 nanostructures as efficient bifunctional electrocatalyst for overall water splitting. J. Power Sources 2019, 416, 95–103. [Google Scholar] [CrossRef]
  16. Da Silva, G.C.; Mayrhofer, K.J.J.; Ticianelli, E.A.; Cherevko, S. Dissolution Stability: The Major Challenge in the Regenerative Fuel Cells Bifunctional Catalysis. J. Electrochem. Soc. 2018, 165, F1376–F1384. [Google Scholar] [CrossRef]
  17. Zhai, P.; Zhang, Y.; Wu, Y.; Gao, J.; Zhang, B.; Cao, S.; Zhang, Y.; Li, Z.; Sun, L.; Hou, J. Engineering active sites on hierarchical transition bimetal oxides/sulfides heterostructure array enabling robust overall water splitting. Nat. Commun. 2020, 11, 5462. [Google Scholar] [CrossRef]
  18. Jung, H.-Y.; Park, S.; Popov, B.N. Electrochemical studies of an unsupported Ptir electrocatalyst as a bifunctional oxygen electrode in a unitized regenerative fuel cell. J. Power Sources 2009, 191, 357–361. [Google Scholar] [CrossRef]
  19. Suryanto, B.H.R.; Wang, Y.; Hocking, R.K.; Adamson, W.; Zhao, C. Overall electrochemical splitting of water at the heterogeneous interface of nickel and iron oxide. Nat. Commun. 2019, 10, 5599. [Google Scholar] [CrossRef]
  20. Hosseini, H.; Roushani, M. Rational design of hollow core-double shells hybrid nanoboxes and nanopipes composed of hierarchical Cu-Ni-Co selenides anchored on nitrogen-doped carbon skeletons as efficient and stable bifunctional electrocatalysts for overall water splitting. Chem. Eng. J. 2020, 402, 126174. [Google Scholar] [CrossRef]
  21. Petala, A.; Noe, A.; Frontistis, Z.; Drivas, C.; Kennou, S.; Mantzavinos, D.; Kondarides, D.I. Synthesis and characterization of CoOx/BiVO4 photocatalysts for the degradation of propyl paraben. J. Hazard. Mater. 2019, 372, 52–60. [Google Scholar] [CrossRef] [PubMed]
  22. Babu, P.; Mohanty, S.; Naik, B.; Parida, K. Serendipitous Assembly of Mixed Phase BiVO4 on B-Doped g-C3N4: An Appropriate p–n Heterojunction for Photocatalytic O2 evolution and Cr(VI) reduction. Inorg. Chem. 2019, 58, 12480–12491. [Google Scholar] [CrossRef] [PubMed]
  23. Jin, S.; Ma, X.; Pan, J.; Zhu, C.; Saji, S.E.; Hu, J.; Xu, X.; Sun, L.; Yin, Z. Oxygen vacancies activating surface reactivity to favor charge separation and transfer in nanoporous BiVO4 photoanodes. Appl. Catal. B Environ. 2021, 281, 119477. [Google Scholar] [CrossRef]
  24. Gaikwad, M.; Ghorpade, U.; Suryawanshi, U.; Kumar, P.; Jang, S.; Jang, S.; Tran, L.; Lee, J.; Bae, H.; Shin, S.; et al. Rapid Synthesis of Ultrathin Ni:FeOOH with In Situ-Induced Oxygen Vacancies for Enhanced Water Oxidation Activity and Stability of BiVO4 Photoanodes. ACS Appl. Mater. Interfaces 2023, 15, 21123–21133. [Google Scholar] [CrossRef] [PubMed]
  25. Baral, B.; Parida, K. {040/110} Facet Isotype Heterojunctions with Monoclinic Scheelite BiVO4. Inorg. Chem. 2020, 59, 10328–10342. [Google Scholar] [CrossRef]
  26. Tan, H.L.; Tahini, H.A.; Wen, X.; Wong, R.J.; Tan, X.; Iwase, A.; Kudo, A.; Amal, R.; Smith, S.C.; Ng, Y.H. Interfacing BiVO4 with Reduced Graphene Oxide for Enhanced Photoactivity: A Tale of Facet Dependence of Electron Shuttling. Small 2016, 12, 5295–5302. [Google Scholar] [CrossRef] [PubMed]
  27. Tan, H.L.; Wen, X.; Amal, R.; Ng, Y.H. BiVO4 {010} and {110} Relative Exposure Extent: Governing Factor of Surface Charge Population and Photocatalytic Activity. J. Phys. Chem. Lett. 2016, 7, 1400–1405. [Google Scholar] [CrossRef] [PubMed]
  28. Murugan, C.; Pandikumar, A. Reinforcement of Visible-Light Harvesting and Charge-Transfer Dynamics of BiVO4 Photoanode via Formation of p−n Heterojunction with CuO for Efficient Photoelectrocatalytic Water Splitting. ACS Appl. Energy Mater. 2022, 5, 6618–6632. [Google Scholar] [CrossRef]
  29. Mary, A.S.; Murugan, C.; Murugan, P.; Pandikumar, A. Unravelling the Superior Photoelectrochemical Water Oxidation Performance of the Al-Incorporated CoOOH Cocatalyst-Loaded BiVO4 Photoanode. ACS Sustain. Chem. Eng. 2023, 11, 13656–13667. [Google Scholar] [CrossRef]
  30. Murugan, C.; Mary, A.S.; Pandikumar, A. Fabrication and Performance Validation of BiVO4 Photoanode-Based Prototype Photoelectrochemical Cells with Different Sizes and Reducing the Photocurrent Density Loss with Different Conductive Patterns. Ind. Eng. Chem. Res. 2024, 63, 4329–4337. [Google Scholar] [CrossRef]
  31. Mary, A.S.; Murugan, C.; Mahendiran, D.; Murugan, P.; Pandikumar, A. Investigation of Mn incorporation into NiOOH electrocatalyst loaded on BiVO4 photoanode for enhanced photoelectrochemical water splitting: Experimental and theoretical approach. Mater. Today Energy 2024, 41, 101541. [Google Scholar] [CrossRef]
  32. Murugan, C.; Mary, A.S.; Velmurugan, R.; Subramanian, B.; Murugan, P.; Pandikumar, A. Investigating the interfacial charge transfer between electrodeposited BiVO4 and pulsed laser-deposited Co3O4 p-n junction photoanode in photoelectrocatalytic water splitting. Chem. Eng. J. 2024, 483, 149104. [Google Scholar] [CrossRef]
  33. Wang, W.; Strohbeen, P.J.; Lee, D.; Zhou, C.; Kawasaki, J.K.; Choi, K.-S.; Liu, M.; Galli, G. The Role of Surface Oxygen Vacancies in BiVO4. Chem. Mater. 2020, 32, 2899–2909. [Google Scholar] [CrossRef]
  34. Li, H.; Yu, H.; Quan, X.; Chen, S.; Zhao, H. Improved Photocatalytic Performance of Heterojunction by Controlling the Contact Facet: High Electron Transfer Capacity between TiO2 and the {110} Facet of BiVO4 Caused by Suitable Energy Band Alignment. Adv. Funct. Mater. 2015, 25, 3074–3080. [Google Scholar] [CrossRef]
  35. Wang, Y.; Tan, G.; Li, B.; Dang, M.; Lv, L.; Wang, M.; Zhang, D.; Ren, H.; Xia, A. Enhanced NIR photocatalytic of Ag-RGO@{010}BiVO4/rgo@{110} BiVO4 photocatalysts induced by resonance effect of transverse electric of RGO and transverse magnetic of Ag. Appl. Surf. Sci. 2019, 489, 1–12. [Google Scholar] [CrossRef]
  36. Qiu, W.; Xiao, S.; Ke, J.; Wang, Z.; Tang, S.; Zhang, K.; Qian, W.; Huang, Y.; Huang, D.; Tong, Y.; et al. Freeing the Polarons to Facilitate Charge Transport in BiVO4 from Oxygen Vacancies with an Oxidative 2D Precursor. Angew. Chem. Int. Ed. 2019, 58, 19087–19095. [Google Scholar] [CrossRef] [PubMed]
  37. Pan, J.; Ma, X.; Zhang, W.; Hu, J. Enhancing the photocatalytic hydrogen production activity of BiVO4 [110] facets using oxygen vacancies. RSC Adv. 2021, 12, 540–545. [Google Scholar] [CrossRef]
  38. Blöchl, P.E. Projector augmented-wave method. Phys. Rev. B. 1994, 50, 17953–17979. [Google Scholar] [CrossRef]
  39. Perdew, J.P.; Burke, K.; Ernzerhof, M. Generalized Gradient Approximation Made Simple. Phys. Rev. Lett. 1996, 77, 3865–3868. [Google Scholar] [CrossRef]
  40. Kresse, G.; Hafner, J. Ab initio molecular-dynamics simulation of the liquid-metal-amorphous-semiconductor transition in germanium. Phys. Rev. B 1994, 49, 14251–14269. [Google Scholar] [CrossRef]
  41. Kresse, G.; Furthmüller, J. Efficiency of ab-initio total energy calculations for metals and semiconductors using a plane-wave basis set. Comput. Mater. Sci. 1996, 6, 15–50. [Google Scholar] [CrossRef]
  42. Thalluri, S.M.; Martinez Suarez, C.; Hussain, M.; Hernandez, S.; Virga, A.; Saracco, G.; Russo, N. Evaluation of the Parameters Affecting the Visible-Light-Induced Photocatalytic Activity of Monoclinic BiVO4 for Water Oxidation. Indust. Engineer. Chem. Res. 2013, 52, 17414–17418. [Google Scholar] [CrossRef]
  43. Anke, B.; Rohloff, M.; Willinger, M.G.; Hetaba, W.; Fischer, A.; Lerch, M. Improved photoelectrochemical performance of bismuth vanadate by partial O/F-substitution. Solid State Sci. 2017, 63, 1–8. [Google Scholar] [CrossRef]
  44. Lardhi, S.; Cavallo, L.; Harb, M. Significant Impact of Exposed Facets on the BiVO4 Material Performance for Photocatalytic Water Splitting Reactions. J. Phys. Chem. Lett. 2020, 11, 5497–5503. [Google Scholar] [CrossRef]
  45. Pan, J.; Wang, R.; Xu, X.; Hu, J.; Ma, L. Transition metal doping activated basal-plane catalytic activity of two-dimensional 1T’-ReS2 for hydrogen evolution reaction: A first-principles calculation study. Nanoscale 2019, 11, 10402–10409. [Google Scholar] [CrossRef] [PubMed]
  46. Hinnemann, B.; Moses, P.G.; Bonde, J.; Jorgensen, K.P.; Nielsen, J.H.; Horch, S.; Chorkendorff, I.; Norskov, J.K. Biomimetic Hydrogen Evolution:  MoS2 Nanoparticles as Catalyst for Hydrogen Evolution. J. Am. Chem. Soc. 2005, 127, 5308–5309. [Google Scholar] [CrossRef] [PubMed]
  47. Hu, J.; Zhao, X.; Chen, W.; Su, H.; Chen, Z. Theoretical Insight into the Mechanism of Photoelectrochemical Oxygen Evolution Reaction on BiVO4 Anode with Oxygen Vacancy. J. Phys. Chem. C 2017, 121, 18702–18709. [Google Scholar] [CrossRef]
  48. Zhang, X.; Yang, Z.; Lu, Z.; Wang, W. Bifunctional CoNx embedded graphene electrocatalysts for OER and ORR: A theoretical evaluation. Carbon 2018, 130, 112–119. [Google Scholar] [CrossRef]
  49. Nørskov, J.K.; Rossmeisl, J.; Logadottir, A.; Lindqvist, L.; Kitchin, J.R.; Bligaard, T.; Jónsson, H. Origin of the overpotential for oxygen reduction at a fuel-cell cathode. J. Phys. Chem. B 2004, 108, 17886–17892. [Google Scholar] [CrossRef]
  50. Liang, Q.; Brocks, G.; Bieberle-Hütter, A. Oxygen evolution reaction (OER) mechanism under alkaline and acidic conditions. J. Phys. Energy 2021, 3, 026001. [Google Scholar] [CrossRef]
  51. Zhou, X.; Dong, H.; Ren, A.M. Exploring the mechanism of water-splitting reaction in NiOx/β-Ga2O3 photocatalysts by first-principles calculations. Phys. Chem. Chem. Phys. 2016, 18, 11111–11119. [Google Scholar] [CrossRef]
Figure 1. Optimized structures, the density of states (DOS), and the work functions compared to the redox potential of water in BiVO4 (110) facets (a) without and (b) with the Ovac.
Figure 1. Optimized structures, the density of states (DOS), and the work functions compared to the redox potential of water in BiVO4 (110) facets (a) without and (b) with the Ovac.
Nanomaterials 14 01270 g001
Figure 2. The total and partial DOS of H adsorption and hydrogen adsorption free energy (ΔGH*) in BiVO4 (110) facets (a) without and (b) with the Ovac.
Figure 2. The total and partial DOS of H adsorption and hydrogen adsorption free energy (ΔGH*) in BiVO4 (110) facets (a) without and (b) with the Ovac.
Nanomaterials 14 01270 g002
Figure 3. DOS of H2O* HO*, O*, and HOO* adsorbed on BiVO4 (110) facets (a) without and (b) with the Ovac. The insets are charge density difference and the yellow and blue colors represent charge accumulation in depletion.
Figure 3. DOS of H2O* HO*, O*, and HOO* adsorbed on BiVO4 (110) facets (a) without and (b) with the Ovac. The insets are charge density difference and the yellow and blue colors represent charge accumulation in depletion.
Nanomaterials 14 01270 g003
Figure 4. Work functions of H2O*, HO*, O*, and HOO* adsorbed on BiVO4 (110) facets (a) without and (b) with the Ovac.
Figure 4. Work functions of H2O*, HO*, O*, and HOO* adsorbed on BiVO4 (110) facets (a) without and (b) with the Ovac.
Nanomaterials 14 01270 g004
Figure 5. Free energy diagrams for the four steps of the OER on the BiVO4 (110) facets without (a) and with (b) the Ovac at U = 0 eV, pH = 0, and T = 298 K.
Figure 5. Free energy diagrams for the four steps of the OER on the BiVO4 (110) facets without (a) and with (b) the Ovac at U = 0 eV, pH = 0, and T = 298 K.
Nanomaterials 14 01270 g005
Figure 6. The Bader charge of active site Bi at each electron step for the OER in the BiVO4 (110) facets with and without the Ovac.
Figure 6. The Bader charge of active site Bi at each electron step for the OER in the BiVO4 (110) facets with and without the Ovac.
Nanomaterials 14 01270 g006
Figure 7. Optimized structure and free energy diagrams for the four steps of the OER on the BiVO4 (110) facets with different Ovac sites and concentrations of (a) 3.13%, (b) 6.25%, and (c) 9.38%.
Figure 7. Optimized structure and free energy diagrams for the four steps of the OER on the BiVO4 (110) facets with different Ovac sites and concentrations of (a) 3.13%, (b) 6.25%, and (c) 9.38%.
Nanomaterials 14 01270 g007
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Fu, H.; Qi, Q.; Li, Y.; Pan, J.; Zhong, C. Oxygen-Vacancy-Induced Enhancement of BiVO4 Bifunctional Photoelectrochemical Activity for Overall Water Splitting. Nanomaterials 2024, 14, 1270. https://doi.org/10.3390/nano14151270

AMA Style

Fu H, Qi Q, Li Y, Pan J, Zhong C. Oxygen-Vacancy-Induced Enhancement of BiVO4 Bifunctional Photoelectrochemical Activity for Overall Water Splitting. Nanomaterials. 2024; 14(15):1270. https://doi.org/10.3390/nano14151270

Chicago/Turabian Style

Fu, Huailiang, Qingxiu Qi, Yushu Li, Jing Pan, and Chonggui Zhong. 2024. "Oxygen-Vacancy-Induced Enhancement of BiVO4 Bifunctional Photoelectrochemical Activity for Overall Water Splitting" Nanomaterials 14, no. 15: 1270. https://doi.org/10.3390/nano14151270

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop