Next Article in Journal
ZnO Matrices as a Platform for Tunable Localized Surface Plasmon Resonances of Silver Nanoparticles
Previous Article in Journal
Effect of Carbon-Doped Cu(Ni) Alloy Film for Barrierless Copper Interconnect
Previous Article in Special Issue
Preparation and Characterization of Polymer-Based Electrospun Nanofibers for Flexible Electronic Applications
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

First-Principles Study of Electronic Structure and Optical Properties of Ni-Doped Bi4O5Br2

1
College of Mathematics & Physics, Weinan Normal University, Weinan 714000, China
2
School of Physics and Electronic Information, Yan’an University, Yan’an 716000, China
3
Network Information Center, Yan’an University, Yan’an 716000, China
4
Science and Technology on Aerospace Chemical Power Laboratory, Xiangyang 441003, China
5
Hubei Institute of Aerospace Chemotechnology, Xiangyang 441003, China
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Coatings 2024, 14(1), 67; https://doi.org/10.3390/coatings14010067
Submission received: 13 December 2023 / Revised: 25 December 2023 / Accepted: 2 January 2024 / Published: 3 January 2024
(This article belongs to the Special Issue Advanced Materials for Electrocatalysis and Energy Storage)

Abstract

:
In this study, we comprehensively explored the electronic structure and optical properties of Ni-doped Bi4O5Br2 through first-principles computational calculations. By calculating its electronic structure and band characteristics, we investigated the impact of Ni doping on the photocatalytic performance of Bi4O5Br2. The computational results indicated that Ni doping significantly altered the band structure of Bi4O5Br2, leading to a reduction in the band gap width. The band gap for undoped Bi4O5Br2 was 2.151 eV, whereas the Ni-doped system exhibited a smaller band gap, directly indicating its enhanced visible light absorption capacity and facilitating the effective separation of photo-generated electron–hole pairs. Through analysis of 2D charge density maps, we observed changes in chemical bonding induced by Ni doping. The shortening of Ni-O bonds suggested increased bond strength, consistent with the observed reduction in cell volume. These findings provide a theoretical foundation for understanding the mechanisms behind the enhanced photocatalytic hydrogen production performance in Ni-doped Bi4O5Br2, offering valuable insights for the design and optimization of highly efficient photocatalytic materials.

1. Introduction

In 1976, Japanese scientist Fujishima made a groundbreaking discovery that TiO2 could catalyze solar energy conversion through photocatalysis [1]. This discovery sparked widespread interest among scientists, leading to significant advancements in energy and environmental management [2,3]. Photocatalysis has demonstrated immense potential in various applications, including water splitting to produce hydrogen [4], CO2 reduction into organic compounds like methane, methanol, and formic acid [5], and ammonia synthesis from N2 [6], providing a clean and sustainable alternative to conventional methods.
Furthermore, photocatalysis plays a crucial role in eco-friendly water pollution control, preventing secondary pollution [7]. Various materials have been explored as photocatalysts, including common sulfide photocatalysts like CdS [8], semiconductor oxide materials such as TiO2 and ZnO [9,10], as well as materials like g-C3N4 and bismuth oxide semiconductors (BixOyHz, where H=Cl, Br, I) [11,12]. Additionally, research has extended to encompass metal-organic framework materials and covalent organic framework materials [13,14,15]. These materials have encountered limitations related to band gaps, light absorption ranges, and electron–hole recombination during photocatalysis [16,17]. Researchers have employed diverse strategies to address these challenges, including constructing heterojunctions, loading materials to inhibit electron–hole recombination, and doping to modify the energy band structure [18,19,20,21,22]. Among these approaches, doping stands out as the simplest and most effective method [23,24,25].
Bismuth oxybromide (Bi4O5Br2), as a ternary (V-VI-VII) semiconductor with the molecular formula BiOX (X = Cl, Br, I), has garnered significant attention in the field of photocatalysis due to its suitable band gap, unique two-dimensional (2D) layered structure, and broad spectrum light responsiveness [26,27,28,29,30,31]. Among these BiOX materials, BiOBr, with its appropriate band structure, has been a focal point of interest, yet there is still room for improvement in its catalytic activity and light absorption range [32,33,34,35,36]. In particular, Bi4O5Br2 nanostructures have drawn our attention due to their broader visible light absorption edge and negative conduction band position when compared to pure BiOBr materials [37,38,39]. However, for the practical application of Bi4O5Br2 in engineering, there are still some limitations to address. For instance, it exhibits a relatively high charge carrier recombination rate and a larger band gap, which restricts its photocatalytic efficiency and also limits its absorption of visible light [40].
Despite the abundance of research on bismuth oxybromide (Bi4O5Br2) photocatalytic materials and doping modifications, there has been relatively limited systematic theoretical investigation into their electronic structure and optical properties [41,42,43,44]. Recently, in the work by Wang et al., the enhancement of photocatalytic performance has been achieved by doping Ni into BiOBr to increase the unit cell dipole moment and induce spin polarization, thereby augmenting the built-in electric field [45]. Therefore, this study employs density functional theory-based pseudopotential plane-wave methods to systematically calculate the electronic structure, density of states, complex dielectric function, absorption spectra, and other properties of Ni-doped Bi4O5Br2. The aim is to theoretically investigate how Ni doping enhances the photocatalytic performance of Bi4O5Br2. This research provides crucial theoretical insights for the design and application of similar studies in the future.

2. Computational Methods

Bi4O5Br2 belongs to the P21 space group with a = 10.94 Å, b = 5.65 Å, c = 14.60 Å, α = γ = 90°, β = 98.02° [46]. To calculate exchange-correlation energy, we applied the Perdew-Burke-Ernzerh (PBE) generalized gradient approximation (GGA) based on density generalization theory using the CASTEP code [47,48,49]. Geometric optimization employs the Broyden–Fletcher–Goldfarb–Shanno (BFGS) method, and core electron interactions are computed with OTFG ultrasoft pseudopotential [50,51]. For geometric optimization and property calculations, we used a 6 × 6 × 2 k-point mesh and an energy cut-off of 580 eV. During atomic relaxation, energy convergence tolerance was set to no more than 1 × 10−5 eV/atom, and atomic forces were limited to less than 0.03 eV/Å, with maximum stress and displacement capped at 0.05 GPa and 0.001 Å.
All computational models were constructed using Bi4O5Br2 (Figure 1a) original cells, which contained 16 Bi atoms, 20 O atoms, and 8 Br atoms. The model of the single-doped system is to replace one Bi atom with Ni. In our calculations, we only considered nickel replacement doping Bi4O5Br2 (Figure 1b) because the radius and valence states of nickel atoms favor replacement doping rather than gap doping [52].

3. Results and Discussion

3.1. Structure Optimization

After Ni replaces one Bi atom in the Bi4O5Br2 primary cell, the lattice constants obtained were a = 5.74, b = 10.99, and c = 14.73 nm, which are 0.05 nm smaller than those obtained by calculating Bi4O5Br2 for a, 0.15 nm smaller than for b, and 0.27 nm larger than for c.
All the parameters of the two structures are listed in Table 1, and the main reason for the analysis is that the radius of the Ni atoms is smaller than that of the Bi atoms and that Ni is less electronegative, forming stronger covalent bonds with O after replacing Bi atoms. This is mainly due to the fact that the radius of Ni atom is smaller than that of Bi, and Ni is less electronegative and forms stronger covalent bonds with O after the replacement of Bi atoms. The calculated Ni-O bond length is shorter than the Bi-O bond, which increases the lattice constant. Chen et al. predicted through DFT calculations that by adjusting the reaction solution pH during the solvent–thermal synthesis process, the crystal lattice parameter b of Bi2MoO6 could be shortened, inducing a larger internal polarization [53]. Similarly, in our study, we achieved a reduction in lattice parameters through Ni doping, which significantly affects the distribution of electron density, thereby inducing the distribution of electron–hole pairs and indirectly improving the photocatalytic performance.
To assess the comparative stability of the analyzed systems, we computed their doped formation energies using the following formula:
E f = E ( N i B i 4 O 5 B r 2 ) E ( pure B i 4 O 5 B r 2 ) + n B i μ B i n N i μ N i
where;
-
E ( N i B i 4 O 5 B r 2 ) is the total energy of the doped Bi4O5Br2,
-
E ( pure B i 4 O 5 B r 2 ) is the total energy of Bi4O5Br2 without doping,
-
nBi and nNi, represent the quantities of added or removed atoms of Bi and Ni, respectively, in the assembly of various cells,
-
and μBi and μNi are the chemical potentials of, respectively; determined through DFT calculations as illustrated in Figure 1c. Each of μBi and μNi corresponds to the energy per atom in their respective bulk crystals, μBi and μNi.
Simultaneously, we computed doping formation energies for Ni separately (refer to Table 2). The lower the doping formation energy, the greater the stability of the resulting structure. Therefore, the structure with the eighth site is deemed optimal.

3.2. Electronic Structures

The depicted energy band structures in Figure 2 illustrate the characteristics of both undoped and Ni-doped materials, with the Fermi energy reference point set at 0 eV. It is important to acknowledge that, in the realm of density functional theory (DFT), the GGA-PBE method is renowned for its limitations, particularly in its tendency to underestimate the material’s band gap when compared to experimental observations. Nevertheless, despite this minor drawback, the obtained band gap remains adequate for conducting a comprehensive exploration of electronic states and investigating optical properties in subsequent analyses.
Furthermore, it is worth noting that the computed band gap for undoped Bi4O5Br2, as depicted in Figure 2a, stands at 2.151 eV, a mere 0.17 eV deviation from the experimentally measured band gap width. While the predictions of PBE are not entirely accurate, it enables a faster prediction of energy trends following structural changes without impacting the analysis of band structures and electronic configurations [54]. As evident in Figure 2, there is a discernible decline in band gap energy as we transition from the undoped system to the Ni-doped system. This trend indirectly implies a gradual enhancement of the material’s light absorption capability. In a recent work, Mao et al., using UV–vis diffuse reflectance spectra and density function theory (DFT) calculations, indicated that Bi4O5Br2 possessed stronger visible light adsorption, which is beneficial to effectively activate molecular oxygen to produce [55]. This implies that the incorporation of Ni can significantly enhance the light absorption capability of Bi4O5Br2, promoting the transfer of electrons and holes as well as the generation of superoxide radicals. Combining these improvements in the catalyst preparation process, the performance of the photocatalyst can be greatly elevated.
To delve deeper into the examination of doping’s impact on electronic properties, we carried out calculations for the partial density of states (PDOS) and electron density distributions around the Fermi energy level for both undoped and doped Bi4O5Br2, as illustrated in Figure 3. Regarding undoped Bi4O5Br2, the partial density of states (PDOS) is presented in Figure 3a. In this context, it is noteworthy that the valence band maxima (VBM) primarily originate from the 2p orbitals of oxygen (O) and the 4p orbitals of bromine (Br), while the conduction band minima (CBM) are predominantly influenced by the 2p orbitals of oxygen (O) and the 6p orbitals of bismuth (Bi).
In the case of Ni-doped Bi4O5Br2, as depicted in Figure 3b, the valence band maximum (VBM) primarily arises from the O-2p orbitals with contributions from Ni-3d orbitals. Conversely, the conduction band minimum (CBM) is predominantly governed by the Bi-6p orbitals. Notably, the introduction of Ni doping, with Ni-3d peaks occurring within the energy range of 0–2 eV, imparts a semi-metallic character to Bi4O5Br2. After Ni doping, a sharp peak in the vicinity of 6 eV is observed in the O s and p orbitals. This is likely due to the formation of a stable chemical bond between Ni and O, resulting in a significant electron density near this energy. This sharp peak may have a crucial impact on electron transport and photocatalytic performance of the material. Additionally, an enhanced density of states (DOS) is observed in the range of 7–10 eV for Bi p orbitals and Br p orbitals, attributed to the contribution of Ni d orbitals. Interaction between Ni d orbitals and Bi/Br likely induces an increase in the electron density of these orbitals. This interaction may lead to the formation of new electronic states, influencing the electronic structure of the catalyst. Consequently, this alteration reduces the energy gap between the valence band and conduction band, resulting in a redshift of the optical absorption edge. This, in turn, enhances the photocatalytic efficiency of the doped system.
In Figure 4a–c,f–h, we present the 2D charge density maps for Bi4O5Br2 and Ni-doped Bi4O5Br2, respectively. Upon analyzing various cross-sections, it becomes evident that the introduction of Ni in place of Bi (Figure 4d,e) results in a reduction in bond length, indicating that the Ni-O bond is stronger compared to the Bi-O bond. This observation aligns with the findings in Table 1, which indicate a decrease in cell volume following Ni atom doping. As illustrated in Figure 4, the differences in bonding properties between the atoms in the undoped and doped cases are relatively subtle. Furthermore, as demonstrated in Figure 4i, we explored the impact of Ni doping on the electron transfer properties of the Bi4O5Br2 samples by utilizing the charge density difference calculated through DFT. In this representation, the yellow color signifies a gain of electrons, while the blue color indicates electron loss.
It is noteworthy that the difference in charge density highlights an electron-rich environment around the surface of Br atoms. Observing Figure 4i allows for us to conclude that, during the photocatalytic process, the built-in electric field formed after Ni doping suppresses the recombination of electrons (e) and holes (h+), thereby enhancing charge separation efficiency. This enables sustained involvement of photo-generated electrons in reduction reactions, while photo-generated holes participate in oxidation reactions, ultimately improving the photocatalytic performance.

3.3. Optical Properties

Ni doping introduces impurity energy levels that markedly enhance light absorption. The simulated absorption spectra of Bi4O5Br2 and Ni-doped Bi4O5Br2 are depicted in Figure 5a. The imaginary part of ε2(ω) related to the optical absorption of the photocatalytic materials is shown in Figure 5b. The absorption coefficient is expressed as follows:
α ( ω ) = ( 2 ) ω ε 1 ( ω ) 2 + ε 2 ( ω ) 2 ε 1 ( ω ) 1 2
where ε1(ω) and ε2(ω) are the real and imaginary parts of the dielectric function, respectively, and both depend on the optical frequency (ω). When the dielectric function ε2 is not zero, it can be determined that the absorption coefficient is positively correlated with it. Doping induces a redshift in absorption peaks, expanding the absorption range. In the low-energy region, Ni doping significantly enhances absorption, favorably influencing photocatalytic performance.
The light absorption of Bi4O5Br2 in the visible range (1.6–3.1 eV) was poor. Ni doping resulted in a significant enhancement of the light absorption of Bi4O5Br2 in the visible range and extended the absorption range from the visible region to the infrared region. It is proved that Ni doping is favorable for the photocatalytic activity enhancement of Bi4O5Br2.

3.4. Photocatalytic Properties

In a photocatalytic reaction, semiconductor materials initially absorb the energy of photons, which must exceed or match the material’s band gap energy level (Eg). This photon-induced excitation results in the generation of photoexcited electron–hole pairs within the semiconductor. Photoexcited electrons transition to the conduction band, attaining higher energy levels, while photoexcited holes remain in the valence band. Subsequently, these photoexcited charge carriers (electrons and holes) propagate through the semiconductor material, either driven by an applied electric field or moving through diffusion toward the material’s surface. This step is crucial as it facilitates the transport of photoexcited carriers to the catalytic reaction sites. However, during their migration, photoexcited electrons and holes may occasionally recombine, releasing previously absorbed energy, typically in the form of heat or light. Upon reaching the surface of the semiconductor and the catalytic sites, photoexcited electrons and holes participate in the catalytic reaction. Typically, photoexcited electrons are involved in reduction reactions, while photoexcited holes take part in oxidation reactions. This surface-driven redox chemistry ultimately initiates and sustains the catalytic reactions. These catalytic reactions encompass a broad range of processes, including gas conversion, water decomposition, organic compound degradation, and more. In these reactions, chemical energy is usually released, leading to the production of desired products.
Hence, we have derived the CBM and VBM of Bi4O5Br2 through a synergistic approach that combines theoretical insights with DFT calculations. This comprehensive approach enhances our ability to make more accurate predictions regarding Bi4O5Br2 photocatalytic performance. The CBM and VBM potentials of undoped Bi4O5Br2 are empirically calculated using the following formulas [56]:
E C B M = 1 2 E g + χ B i 4 O 5 B r 2 + E 0
E V B M = + 1 2 E g + χ B i 4 O 5 B r 2 + E 0
χ B i 4 O 5 B r 2 = ( χ 4 B i χ 5 O χ 2 B r ) 1 11
where Eg is the band gap energy, χBi4O5Br2 is the absolute electronegativity of Bi4O5Br2. χBi, χBr and χO are the Pearson absolute electronegativity of, respectively. According to the literature: χBi = 4.69, χBr = 7.59 and χO = 7.54 [57]. The parameter E0 serves as the scale factor that connects the reference redox level to the absolute vacuum scale [56]. The obtained CBM of Bi4O5Br2 is 0.77 and VBM is 2.92. While for the Bi4O5Br2 after Ni doping the CBM and VBM are 0.89 and 2.81, respectively. In the case of Ni-doped Bi4O5Br2, the CBM is shifted downward by 0.12 eV with respect to Bi4O5Br2, thus greatly improving the photoreduction capacity. Furthermore, the VBM values of the doped system span the oxidation-reduction potential of hydroxyl radicals. This is consistent with the analysis in the electronic structure section, where changes in the valence band and conduction band are evident in the generation of free radical groups. This can significantly improve the catalytic performance of the catalyst.
Therefore, the mechanism underlying the enhanced photocatalytic activity can be elucidated as follows. Firstly, Ni introduces impurity energy levels with both upper and lower spin in the bandgap of Bi4O5Br2 after doping, altering the electron transition pathways and significantly reducing the bandgap of Bi4O5Br2. Secondly, the optical absorption performance of the Ni-Bi4O5Br2 composite material not only exhibits a pronounced redshift but also shows enhanced absorption intensity. Thus, this catalysis even holds potential for infrared reactions. In comparison to pristine Bi4O5Br2, these two advantages can significantly increase the excitation rate of photo-generated electron–hole pairs. The effective spatial separation of oxidation and reduction centers markedly suppresses the recombination of electron–hole pairs, thereby improving charge carrier utilization. The combined effect of these factors endows the Ni-Bi4O5Br2 composite material with outstanding photocatalytic performance.

4. Conclusions

In order to explore the physical origin of the experimentally observed doping enhancement on the photocatalytic performance of Bi4O5Br2, we conducted a comprehensive study of the electronic structure and optical properties of Ni-doped Bi4O5Br2 through first-principles computational calculations. The computational results indicate that Ni doping significantly alters the band structure of Bi4O5Br2. The band gap for undoped Bi4O5Br2 is 2.151 eV, whereas the Ni-doped system exhibits a smaller band gap, directly indicating its enhanced visible light absorption capacity. In comparison to the undoped system, Ni doping leads to more efficient separation of electrons and holes, further promoting the photocatalytic performance. Through the analysis of 2D charge density maps, we observed changes in chemical bonding induced by Ni doping. The shortening of Ni-O bonds suggests increased bond strength, consistent with the observed reduction in cell volume.
Furthermore, our analysis using DFT-computed charge density difference maps revealed that Ni doping alters the electron transfer and changes the electron-rich environment around Br atoms on the surface. We can conclude that, during the photocatalytic process, the built-in electric field formed after Ni doping suppresses the recombination of electrons (e) and holes (h+), thereby enhancing charge separation efficiency. In terms of light absorption, Ni doping significantly enhances the absorption of Bi4O5Br2 in the visible range, extending the absorption range into the infrared region. This enhancement is not only theoretically supported but also quantitatively contributes to the improved photocatalytic activity of Bi4O5Br2. This study provides concrete guidance for the design of efficient photocatalytic materials and underscores the effectiveness of Ni doping as a strategy to enhance photocatalytic performance.

Author Contributions

F.Z. and X.L. supervised the research; H.S. (Hong Sheng) and Y.L. designed the experimental program; H.S. (Hong Sheng) and Y.L. completed the experimental preparation and wrote the first draft; F.Z. provided funding acquisition and revised data and papers; X.Z., S.X. (Shiheng Xin), G.L., Q.W., S.X. (Suqin Xue), X.W., T.S. and H.S. (Hui Shi) et al. analyzed and confirmed the data; All authors participated in analysis of the experimental data and discussions of the results, as well as editing the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Natural Science Foundation of China (No: 62264015), 2023 Innovation and Entrepreneurship for Undergraduates (2023107 19040).

Data Availability Statement

Data are contained within the article.

Conflicts of Interest

There are no conflicts to declare.

References

  1. Fujishima, A.; Honda, K. Electrochemical Photolysis of Water at a Semiconductor Electrode. Nature 1972, 238, 37–38. [Google Scholar] [CrossRef] [PubMed]
  2. Zou, Z.; Ye, J.; Sayama, K.; Arakawa, H. Direct splitting of water under visible light irradiation with an oxide semiconductor photocatalyst. Nature 2001, 414, 625–627. [Google Scholar] [CrossRef] [PubMed]
  3. Hoffmann, M.R.; Martin, S.T.; Choi, W.; Bahnemann, D.W. Environmental Applications of Semiconductor Photocatalysis. Chem. Rev. 1995, 95, 69–96. [Google Scholar] [CrossRef]
  4. Hisatomi, T.; Kubota, J.; Domen, K. Recent advances in semiconductors for photocatalytic and photoelectrochemical water splitting. Chem. Soc. Rev. 2014, 43, 7520–7535. [Google Scholar] [CrossRef] [PubMed]
  5. Gong, E.; Ali, S.; Hiragond, C.B.; Kim, H.S.; Powar, N.S.; Kim, D.; Kim, H.; In, S.I. Solar fuels: Research and development strategies to accelerate photocatalytic CO2 conversion into hydrocarbon fuels. Energy Environ. Sci. 2022, 15, 880–937. [Google Scholar] [CrossRef]
  6. Garrido-Barros, P.; Derosa, J.; Chalkley, M.J.; Peters, J.C. Tandem electrocatalytic N-2 fixation via proton-coupled electron transfer. Nature 2022, 609, 71–76. [Google Scholar] [CrossRef]
  7. Rueda-Márquez, J.J.; Levchuk, I.; Ibáñez, P.F.; Sillanpää, M. A critical review on application of photocatalysis for toxicity reduction of real wastewaters. J. Clean. Prod. 2020, 258, 120694. [Google Scholar] [CrossRef]
  8. Cheng, L.; Xiang, Q.; Liao, Y.; Zhang, H. CdS-Based photocatalysts. Energy Environ. Sci. 2018, 11, 1362–1391. [Google Scholar] [CrossRef]
  9. Nakata, K.; Ochiai, T.; Murakami, T.; Fujishima, A. Photoenergy conversion with TiO2 photocatalysis: New materials and recent applications. Electrochim. Acta 2012, 84, 103–111. [Google Scholar] [CrossRef]
  10. Agustianingrum, M.P.; Yoshida, S.; Tsuji, N.; Park, N. Effect of aluminum addition on solid solution strengthening in CoCrNi medium-entropy alloy. J. Alloys Compd. 2019, 781, 866–872. [Google Scholar] [CrossRef]
  11. Li, J.; Yu, Y.; Zhang, L. Bismuth oxyhalide nanomaterials: Layered structures meet photocatalysis. Nanoscale 2014, 6, 8473–8488. [Google Scholar] [CrossRef] [PubMed]
  12. Jin, X.; Ye, L.; Xie, H.; Chen, G. Bismuth-rich bismuth oxyhalides for environmental and energy photocatalysis. Coord. Chem. Rev. 2017, 349, 84–101. [Google Scholar] [CrossRef]
  13. Qian, Y.; Zhang, F.; Pang, H. A Review of MOFs and Their Composites-Based Photocatalysts: Synthesis and Applications. Adv. Funct. Mater. 2021, 31, 2104231. [Google Scholar] [CrossRef]
  14. Hu, X.-L.; Li, H.-G.; Tan, B.-E. COFs-based Porous Materials for Photocatalytic Applications. Chin. J. Polym. Sci. 2020, 38, 673–684. [Google Scholar] [CrossRef]
  15. Ji, X.-Y.; Sun, K.; Liu, Z.-K.; Liu, X.; Dong, W.; Zuo, X.; Shao, R.; Tao, J. Identification of Dynamic Active Sites Among Cu Species Derived from MOFs@CuPc for Electrocatalytic Nitrate Reduction Reaction to Ammonia. Nano-Micro Lett. 2023, 15, 110. [Google Scholar] [CrossRef]
  16. Zhang, L.W.; Mohamed, H.H.; Dillert, R.; Bahnemann, D. Kinetics and mechanisms of charge transfer processes in photocatalytic systems: A review. J. Photochem. Photobiol. C-Photochem. Rev. 2012, 13, 263–276. [Google Scholar] [CrossRef]
  17. Wang, W.; Tade, M.O.; Shao, Z.P. Nitrogen-doped simple and complex oxides for photocatalysis: A review. Prog. Mater. Sci. 2018, 92, 33–63. [Google Scholar] [CrossRef]
  18. Liu, Y.; Yang, Y.N.; Zhang, B.H.; Deng, D.; Ning, J.; Liu, G.H.; Xue, S.Q.; Zhang, F.C.; Liu, X.H.; Zhang, W.B. Electronic structure and optical properties of CdS/BiOI heterojunction improved by oxygen vacancies. J. Alloys Compd. 2023, 955, 170235. [Google Scholar] [CrossRef]
  19. Ning, J.; Zhang, B.; Siqin, L.; Liu, G.; Wu, Q.; Xue, S.; Shao, T.; Zhang, F.; Zhang, W.; Liu, X. Designing advanced S-scheme CdS QDs/La-Bi2WO6 photocatalysts for efficient degradation of RhB. Exploration 2023, 3, 20230050. [Google Scholar] [CrossRef]
  20. Liu, H.D.; Cheng, M.; Liu, Y.; Wang, J.; Zhang, G.X.; Li, L.; Du, L.; Wang, G.F.; Yang, S.Z.; Wang, X.Y. Single atoms meet metal-organic frameworks: Collaborative efforts for efficient photocatalysis. Energy Environ. Sci. 2022, 15, 3722–3749. [Google Scholar] [CrossRef]
  21. Ma, H.; Zhao, F.; Li, M.; Wang, P.; Fu, Y.; Wang, G.; Liu, X. Construction of hollow binary oxide heterostructures by Ostwald ripening for superior photoelectrochemical removal of reactive brilliant blue KNR dye. Adv. Powder Mater. 2023, 2, 100117. [Google Scholar] [CrossRef]
  22. Zhang, B.; Fang, C.; Ning, J.; Dai, R.; Liu, Y.; Wu, Q.; Zhang, F.; Zhang, W.; Dou, S.; Liu, X. Unusual aliovalent Cd doped γ-Bi2MoO6 nanomaterial for efficient photocatalytic degradation of sulfamethoxazole and rhodamine B under visible light irradiation. Carbon Neutralization 2023, 2, 646–660. [Google Scholar] [CrossRef]
  23. Park, H.; Park, Y.; Kim, W.; Choi, W. Surface modification of TiO2 photocatalyst for environmental applications. J. Photochem. Photobiol. C-Photochem. Rev. 2013, 15, 1–20. [Google Scholar] [CrossRef]
  24. Low, J.X.; Yu, J.G.; Jaroniec, M.; Wageh, S.; Al-Ghamdi, A.A. Heterojunction Photocatalysts. Adv. Mater. 2017, 29, 1601694. [Google Scholar] [CrossRef] [PubMed]
  25. Zhang, B.; Liu, G.; Shi, H.; Wu, Q.; Xue, S.; Shao, T.; Fuchun, Z.; Liu, X. Density Functional Theory Study of Electronic Structure and Optical Properties of Ln3+-Doped γ-Bi2MoO6 (Ln = Gd, Ho, Yb). Crystals 2023, 13, 1158. [Google Scholar] [CrossRef]
  26. Chen, J.; Guan, M.L.; Cai, W.Z.; Guo, J.J.; Xiao, C.; Zhang, G.K. The dominant {001} facet-dependent enhanced visible-light photoactivity of ultrathin BiOBr nanosheets. Phys. Chem. Chem. Phys. 2014, 16, 20909–20914. [Google Scholar] [CrossRef]
  27. Di, J.; Xia, J.X.; Ji, M.X.; Wang, B.; Yin, S.; Xu, H.; Chen, Z.G.; Li, H.M. Carbon Quantum Dots Induced Ultrasmall BiOI Nanosheets with Assembled Hollow Structures for Broad Spectrum Photocatalytic Activity and Mechanism Insight. Langmuir 2016, 32, 2075–2084. [Google Scholar] [CrossRef]
  28. He, R.A.; Cao, S.W.; Zhou, P.; Yu, J.G. Recent advances in visible light Bi-based photocatalysts. Chin. J. Catal. 2014, 35, 989–1007. [Google Scholar] [CrossRef]
  29. Li, H.; Shang, J.; Shi, J.G.; Zhao, K.; Zhang, L.Z. Facet-dependent solar ammonia synthesis of BiOCl nanosheets via a proton-assisted electron transfer pathway. Nanoscale 2016, 8, 1986–1993. [Google Scholar] [CrossRef]
  30. Li, J.; Zhan, G.M.; Yu, Y.; Zhang, L.Z. Superior visible light hydrogen evolution of Janus bilayer junctions via atomic-level charge flow steering. Nat. Commun. 2016, 7, 11480. [Google Scholar] [CrossRef]
  31. Xu, J.; Li, L.; Guo, C.S.; Zhang, Y.; Wang, S.F. Removal of benzotriazole from solution by BiOBr photocatalysis under simulated solar irradiation. Chem. Eng. J. 2013, 221, 230–237. [Google Scholar] [CrossRef]
  32. Cheng, H.F.; Huang, B.B.; Dai, Y. Engineering BiOX (X = Cl, Br, I) nanostructures for highly efficient photocatalytic applications. Nanoscale 2014, 6, 2009–2026. [Google Scholar] [CrossRef] [PubMed]
  33. Ai, Z.H.; Wang, J.L.; Zhang, L.Z. Substrate-dependent photoreactivities of BiOBr nanoplates prepared at different pH values. Chin. J. Catal. 2015, 36, 2145–2154. [Google Scholar] [CrossRef]
  34. Di, J.; Xia, J.X.; Ji, M.X.; Wang, B.; Li, X.W.; Zhang, Q.; Chen, Z.G.; Li, H.M. Nitrogen-Doped Carbon Quantum Dots/BiOBr Ultrathin Nanosheets: In Situ Strong Coupling and Improved Molecular Oxygen Activation Ability under Visible Light Irradiation. Acs Sustain. Chem. Eng. 2016, 4, 136–146. [Google Scholar] [CrossRef]
  35. Di, J.; Xia, J.X.; Ji, M.X.; Wang, B.; Yin, S.; Zhang, Q.; Chen, Z.G.; Li, H.M. Advanced photocatalytic performance of graphene-like BN modified BiOBr flower-like materials for the removal of pollutants and mechanism insight. Appl. Catal. B-Environ. 2016, 183, 254–262. [Google Scholar] [CrossRef]
  36. Li, R.; Gao, X.Y.; Fan, C.M.; Zhang, X.C.; Wang, Y.W.; Wang, Y.F. A facile approach for the tunable fabrication of BiOBr photocatalysts with high activity and stability. Appl. Surf. Sci. 2015, 355, 1075–1082. [Google Scholar] [CrossRef]
  37. Chang, F.; Lei, B.; Yang, C.; Wang, J.Y.; Hu, X.F. Ultra-stable Bi4O5Br2/Bi2S3 n-p heterojunctions induced simultaneous generation of radicals center dot OH and center dot O2− and NO conversion to nitrate/nitrite species with high selectivity under visible light. Chem. Eng. J. 2021, 413, 127443. [Google Scholar] [CrossRef]
  38. Chang, F.; Yang, C.; Wang, J.Y.; Lei, B.; Li, S.J.; Kim, H. Enhanced photocatalytic conversion of NOx with satisfactory selectivity of 3D-2D Bi4O5Br2-GO hierarchical structures via a facile. Sep. Purif. Technol. 2021, 266, 118237. [Google Scholar] [CrossRef]
  39. Wu, Z.H.; Shen, J.; Ma, N.; Li, Z.F.; Wu, M.; Xu, D.F.; Zhang, S.Y.; Feng, W.H.; Zhu, Y.F. Bi4O5Br2 nanosheets with vertical aligned facets for efficient visible-light-driven photodegradation of BPA. Appl. Catal. B-Environ. 2021, 286, 119937. [Google Scholar] [CrossRef]
  40. Li, R.; Liu, J.X.; Zhang, X.F.; Wang, Y.W.; Wang, Y.F.; Zhang, C.M.; Zhang, X.C.; Fan, C.M. Iodide-modified Bi4O5Br2 photocatalyst with tunable conduction band position for efficient visible-light decontamination of pollutants. Chem. Eng. J. 2018, 339, 42–50. [Google Scholar] [CrossRef]
  41. Bai, Y.; Yang, P.; Wang, L.; Yang, B.; Xie, H.; Zhou, Y.; Ye, L. Ultrathin Bi4O5Br2 nanosheets for selective photocatalytic CO2 conversion into CO. Chem. Eng. J. 2019, 360, 473–482. [Google Scholar] [CrossRef]
  42. Zhang, W.-B.; Xiao, X.; Wu, Q.-F.; Fan, Q.; Chen, S.; Yang, W.-X.; Zhang, F.-C. Facile synthesis of novel Mn-doped Bi4O5Br2 for enhanced photocatalytic NO removal activity. J. Alloys Compd. 2020, 826, 154204. [Google Scholar] [CrossRef]
  43. Zhu, G.; Hojamberdiev, M.; Zhang, W.; Taj Ud Din, S.; Joong Kim, Y.; Lee, J.; Yang, W. Enhanced photocatalytic activity of Fe-doped Bi4O5Br2 nanosheets decorated with Au nanoparticles for pollutants removal. Appl. Surf. Sci. 2020, 526, 146760. [Google Scholar] [CrossRef]
  44. Xiao, X.; Lu, M.; Nan, J.; Zuo, X.; Zhang, W.; Liu, S.; Wang, S. Rapid microwave synthesis of I-doped Bi4O5Br2 with significantly enhanced visible-light photocatalysis for degradation of multiple parabens. Appl. Catal. B Environ. 2017, 218, 398–408. [Google Scholar] [CrossRef]
  45. Wang, Y.; Xie, Y.; Yu, S.; Yang, K.; Shao, Y.; Zou, L.; Zhao, B.; Wang, Z.; Ling, Y.; Chen, Y. Ni doping in unit cell of BiOBr to increase dipole moment and induce spin polarization for promoting CO2 photoreduction via enhanced build-in electric field. Appl. Catal. B Environ. 2023, 327, 122420. [Google Scholar] [CrossRef]
  46. Keller, E.; Ketterer, J.; Kramer, V. Crystal structure and twinning of Bi4O5Br2. Z. Fur Krist. 2001, 216, 595–599. [Google Scholar] [CrossRef]
  47. Segall, M.D.; Lindan, P.J.D.; Probert, M.J.; Pickard, C.J.; Hasnip, P.J.; Clark, S.J.; Payne, M.C. First-principles simulation: Ideas, illustrations and the CASTEP code. J. Phys.-Condens. Matter 2002, 14, 2717–2744. [Google Scholar] [CrossRef]
  48. Perdew, J.P.; Burke, K.; Ernzerhof, M. Generalized Gradient Approximation Made Simple. Phys. Rev. Lett. 1996, 77, 3865–3868. [Google Scholar] [CrossRef]
  49. Monkhorst, H.J.; Pack, J.D. Special points for Brillouin-zone integrations. Phys. Rev. B 1976, 13, 5188–5192. [Google Scholar] [CrossRef]
  50. Vanderbilt, D. Soft self-consistent pseudopotentials in a generalized eigenvalue formalism. Phys. Rev. B 1990, 41, 7892–7895. [Google Scholar] [CrossRef]
  51. Fischer, T.H.; Almlöf, J. General methods for geometry and wave function optimization. J. Phys. Chem. 1992, 96, 9768–9774. [Google Scholar] [CrossRef]
  52. Wang, Z.L.; Gu, Y.A.; Zheng, L.X.; Hou, J.W.; Zheng, H.J.; Sun, S.J.; Wang, L.Z. Machine Learning Guided Dopant Selection for Metal Oxide-Based Photoelectrochemical Water Splitting: The Case Study of Fe2O3 and CuO. Adv. Mater. 2022, 34, 2106776. [Google Scholar] [CrossRef] [PubMed]
  53. Chen, Y.; Yang, W.; Gao, S.; Zhu, L.; Sun, C.; Li, Q. Internal Polarization Modulation in Bi2MoO6 for Photocatalytic Performance Enhancement under Visible-Light Illumination. ChemSusChem 2018, 11, 1521–1532. [Google Scholar] [CrossRef] [PubMed]
  54. Stampfl, C.; Van de Walle, C.G. Density-functional calculations for III-V nitrides using the local-density approximation and the generalized gradient approximation. Phys. Rev. B 1999, 59, 5521–5535. [Google Scholar] [CrossRef]
  55. Mao, D.; Ding, S.; Meng, L.; Dai, Y.; Sun, C.; Yang, S.; He, H. One-pot microemulsion-mediated synthesis of Bi-rich Bi4O5Br2 with controllable morphologies and excellent visible-light photocatalytic removal of pollutants. Appl. Catal. B Environ. 2017, 207, 153–165. [Google Scholar] [CrossRef]
  56. Zhang, C.; Jiang, N.; Xu, S.; Li, Z.; Liu, X.; Cheng, T.; Han, A.; Lv, H.; Sun, W.; Hou, Y. Towards high visible light photocatalytic activity in rare earth and N co-doped SrTiO3: A first principles evaluation and prediction. RSC Adv. 2017, 7, 16282–16289. [Google Scholar] [CrossRef]
  57. Pearson, R.G. Absolute electronegativity and hardness: Application to inorganic chemistry. Inorg. Chem. 1988, 27, 734–740. [Google Scholar] [CrossRef]
Figure 1. (a) Bi4O5Br2 (b) Ni-Doped Bi4O5Br2 (c) The structure used to calculate the chemical potential of Bi and Ni.
Figure 1. (a) Bi4O5Br2 (b) Ni-Doped Bi4O5Br2 (c) The structure used to calculate the chemical potential of Bi and Ni.
Coatings 14 00067 g001
Figure 2. Energy band (a) Bi4O5Br2, after considering spin (b) majority spin (c) minority spin.
Figure 2. Energy band (a) Bi4O5Br2, after considering spin (b) majority spin (c) minority spin.
Coatings 14 00067 g002
Figure 3. Total state density (TDOS) and projected state density (PDOS) of (a) Bi4O5Br2 and (b) Ni-doped Bi4O5Br2.
Figure 3. Total state density (TDOS) and projected state density (PDOS) of (a) Bi4O5Br2 and (b) Ni-doped Bi4O5Br2.
Coatings 14 00067 g003
Figure 4. 2D charge-density plots (ac) without Ni and (fh) with Ni doped sections in different directions. (d,e) The change in bond length of Ni-O and Bi-O after replacement doping. (i) 3D charge density map after doping.
Figure 4. 2D charge-density plots (ac) without Ni and (fh) with Ni doped sections in different directions. (d,e) The change in bond length of Ni-O and Bi-O after replacement doping. (i) 3D charge density map after doping.
Coatings 14 00067 g004
Figure 5. Imaginary part of (a) light absorption range (b) dielectric function of Bi4O5Br2 and Ni-Bi4O5Br2.
Figure 5. Imaginary part of (a) light absorption range (b) dielectric function of Bi4O5Br2 and Ni-Bi4O5Br2.
Coatings 14 00067 g005
Table 1. Lattice parameters before and after optimization.
Table 1. Lattice parameters before and after optimization.
abcαβγVolume
Bi4O5Br25.7911.1415.0082.4389.9990.01960.41
Ni-Bi4O5Br25.7410.9914.7382.3591.0587.87950.17
Table 2. Calculated doping formation energy at different locations.
Table 2. Calculated doping formation energy at different locations.
Site12345678
Formation energy/eV1.761.702.422.442.622.342.711.91
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

Sheng, H.; Zhang, X.; Xin, S.; Shi, H.; Liu, G.; Wu, Q.; Xue, S.; Wang, X.; Shao, T.; Liu, Y.; et al. First-Principles Study of Electronic Structure and Optical Properties of Ni-Doped Bi4O5Br2. Coatings 2024, 14, 67. https://doi.org/10.3390/coatings14010067

AMA Style

Sheng H, Zhang X, Xin S, Shi H, Liu G, Wu Q, Xue S, Wang X, Shao T, Liu Y, et al. First-Principles Study of Electronic Structure and Optical Properties of Ni-Doped Bi4O5Br2. Coatings. 2024; 14(1):67. https://doi.org/10.3390/coatings14010067

Chicago/Turabian Style

Sheng, Hong, Xin Zhang, Shiheng Xin, Hui Shi, Gaihui Liu, Qiao Wu, Suqin Xue, Xiaoyang Wang, Tingting Shao, Yang Liu, and et al. 2024. "First-Principles Study of Electronic Structure and Optical Properties of Ni-Doped Bi4O5Br2" Coatings 14, no. 1: 67. https://doi.org/10.3390/coatings14010067

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