Skip to main content
ACS Omega logoLink to ACS Omega
. 2020 Nov 5;5(45):29081–29091. doi: 10.1021/acsomega.0c03685

Hybrid Density Functional Theory Study of Native Defects and Nonmetal (C, N, S, and P) Doping in a Bi2WO6 Photocatalyst

Jihua Zhang , Panluo Deng , Mingsen Deng , Hujun Shen , Zhenzhen Feng , Hang Li ‡,*
PMCID: PMC7675596  PMID: 33225139

Abstract

graphic file with name ao0c03685_0009.jpg

Native defects and nonmetal doping have been shown to be an effective way to optimize the photocatalytic properties of Bi2WO6. However, a detailed understanding of defect physics in Bi2WO6 has been lacking. Here, using the Heyd–Scuseria–Ernzerhof hybrid functional defect calculations, we study the formation energies, electronic structures, and optical properties of native defects and nonmetal element (C, N, S, and P) doping into Bi2WO6. We find that the Bi vacancy (Bivac), O vacancy (Ovac), S doping on the O site (SO), and N doping on the O site (NO) defects in the Bi2WO6 can be stable depending on the Fermi level and chemical potentials. By contrast, the substitution of an O atom by a C or P atom (CO, PO) has high formation energy and is unlikely to form. The calculated electronic structures of the Bivac, Ovac, SO, and NO defects indicate that the band-gap reduction of Ovac2+, Bivac, and SO defects is mainly due to forming shallow impurity levels within the band gap. The calculated absorption coefficients of Ovac2+, Bivac, and SO show strong absorption in the visible light region, which is in good agreement with the experimental results. Hence, Ovac2+, Bivac, and SO defects can improve the adsorption capacity of Bi2WO6, which helps enhance its photocatalytic performance. Our results provide insights into how to enhance the photocatalytic activity of Bi2WO6 for energy and environmental applications through the rational design of defect-controlled synthesis conditions.

1. Introduction

Visible-light-driven photocatalysts are promising candidates for solving the energy crisis and environmental pollution problems because of their high efficiency and clean and renewable characteristics.15 Recently, Bi3+-based oxides such as BiOCl,69 Bi2MoO6,10 and Bi2WO61113 are fascinating and of practical importance for visible-light-driven photocatalysts. Among these materials, Bi2WO6, as an n-type semiconductor with a 2.80 eV13 band gap, is an interesting Aurivillius phase of perovskite layers inserted between Bi2O2 layers. Its potential applications, including solar cells, electrode materials, an excellent photocatalyst for water splitting, and photodegradation of organic pollutants.11,1417

However, the visible-light-responsive photocatalytic application of Bi2WO611,18,19 is limited by rapid recombination of excitonic pairs, ineffective charge separation, and poor visible-light absorption. Various improvement methods have been developed to improve the photocatalytic activity, for example, semiconductor coupling,20 morphology control,21,22 noble metal deposition,23 and metal or nonmetal ion doping.2433 All these methods can improve the photocatalytic activity of Bi2WO6. Nonmetallic element doping is regarded as an efficient approach to enhance the activity. For example, Hoang et al.30 synthesized N-doped Bi2WO6 by using a two-step microwave-assisted and hydrothermal method. It was found that 0.25% (atomic ratio) N-doped Bi2WO6 exhibited the best photocatalytic activity. This is due to the most significant decrease in the recombination rate of photogenerated electron–hole pairs. Fu et al.29 reported that doping Bi2WO6 with 0.5% B displayed the highest photocatalytic activity under visible-light irradiation. This is because the doped B atoms could act as electron traps and facilitate the separation of photogenerated electron–hole pairs. Shang et al.31 demonstrated that doping Bi2WO6 with N could significantly enhance photocatalytic activities. The reasons are that N doped into Bi2WO6 not only broadens the range of light absorption in the visible region but also inhibits the photogenerated electron–hole recombination losses and increases the transfer rate. However, the nature of the promoting effects remains elusive. A complete understanding of the nonmetallic-induced photocatalytic activity of Bi2WO6 is still lacking.

The formation of intrinsic defects (such as Bi vacancy, W vacancy, O vacancy, and so on) may compensate the extrinsic nonmetallic doping into Bi2WO6. Therefore, the intrinsic defects of Bi2WO6 had been studied both theoretically and experimentally. For example, Huo et al.34 showed that the photocatalytic NO oxidation activity was enhanced by introducing oxygen vacancy defects into Bi2WO6. Using the GGA (generalized gradient approximation) + U method, they reported that oxygen vacancies narrow band gaps compared with perfect Bi2WO6. Di et al.15 had engineered bismuth vacancies into Bi2WO6 using a highly efficient template-directed strategy. They found that bismuth vacancies favor water oxidation reactions. Their PBE (Perdew–Burke–Ernzerhof) calculations suggest that bismuth vacancies in Bi2WO6 would induce new energy levels in the band gap, increasing its conductivity.

Theoretically, intrinsic defects and nonmetallic doping in Bi2WO6 have been calculated with the GGA-PBE and DFT + U method.15,28,3436 However, they only focus on neutral impurities and their electronic properties. Note that the formation energy of a neutral defect is independent of the position of the Fermi level. Whether an impurity or defect becomes electrically active or instead remains inactive often depends on the growth conditions (e.g., the presence of native defects) and cannot directly be inferred from a DFT calculation on a neutral defect alone.37 In addition, GGA-PBE yield significantly underestimated band gaps, i.e., 1.85–1.95 eV, compared to the experimental value of 2.80 eV.13 Moreover, the GGA + U correction was only employed for W’s d electrons, whereas all other components were treated within GGA. Hence, quantitatively more reliable calculations are required. As proposed by Heyd, Scuseria, and Ernzerhof (HSE06),38,39 it provides the best overall performance compared to GGA-PBE and GGA + U. More importantly, previous results have shown that the HSE06 hybrid functional is a reliable method to calculate transition levels and formation energies of defects.40

This paper systematically studies the doping properties for both intrinsic and nonmetal (C, N, S, and P) doping of Bi2WO6 and understands how they can change the electronic structure and optical absorption properties using the HSE06 hybrid approach. We calculate the transition energies and formation energies of intrinsic and impurity defects and Fermi-level pinning positions. The Bivac, Ovac, SO, and NO have rather low formation energies. Due to CO and PO having higher formation energies, they are unlikely to form. The calculated electronic structures of Bivac, Ovac, SO, and NO defects implied that Ovac2+, Bivac, and SO defects can narrow the band gap of Bi2WO6, and no recombination centers occurred, which can improve photocatalysis of Bi2WO6. The calculated absorption coefficients of Ovac2+, Bivac, and SO show strong absorption in the visible region. Our results suggest that Ovac2+, Bivac, and SO doping is able to narrow the band gap, and thus these defects are good dopant candidates to tailor the visible light absorption property of Bi2WO6 photocatalysts.

2. Results and Discussion

2.1. Formation Energy

For vacancy calculations, one atom is removed from the supercell, and the remaining atoms are allowed to relax. Here, we consider three types of intrinsic defects in the Bi2WO6 lattice for various possible charge states, including Bivac, Wvac, and Ovac. The substitutional defect XO is created by replacing a lattice O by X (X = C, N, S, and P). The defect formation energies for various charge states in Bi2WO6 are plotted as a function of the electronic Fermi energy for five representative chemical potential conditions with the finite size corrections in Figure 1. As shown in Figure 1, all the defects’ formation energies depend on the growth condition and the Fermi level. The energy of a defect with a charge state is only shown for the range of Fermi levels where the charge state has the lowest energy. In Figure 1, the line’s slope is associated with the most stable charge state of a defect, and the transition levels lie where the slopes change. Moreover, Figure 1 depicts that the formation energies of intrinsic defects and nonmetal-doped (C, N, S, and P) defects in the charge-neutral charge states are constant regardless of the variations in Fermi energy.

Figure 1.

Figure 1

Defect formation energy calculations of Bi2WO6 supercells with various types of point defects in different charge states using the HSE06 method as a function of the Fermi level (referenced to valence band maximum) under (a) C, (b) D, (c) E, (d) F, and (e) M growth conditions (see Figure 5). The energetic value of the Fermi level pinning was indicated by red dashed lines. The slope of the curves indicates the charge state, and the kinks in the curves indicate the transition of charge states. The numbers in the plot indicate the defect charge state; parallel lines imply equal charge states. For each defect, only the charge states that are energetically most favorable at a given Fermi energy are shown.

In Figure 1a,c,e, at chemical potentials C, E, and M, the results show that the dominant native acceptor is Bivac, while the primary native donor is Ovac. The Fermi level is pinned (where the formation energy lines cross the energy overlap of the lowest-energy donor and acceptor) by Bivac and Ovac. However, the position of the Fermi level pinning is different. At chemical potentials C, the Fermi level is pinned to be at ∼1.1 eV above VBM, showing a p-type conductivity. At chemical potential E, as Ovac is positively charged and Bivac is negatively charged for Fermi level within the band gap, these defects self-compensate, and the Fermi level is pinned roughly around the midgap, resulting in the intrinsically insulating property. At chemical potential M, the Fermi level is pinned to be at ∼0.75 eV below CBM within the n-type region. These results imply that it is possible to control the growth condition to obtain the conductivity from p-type to good n-type. At chemical potentials C and E, the formation energy of the Wvac defect is similar to that of OS and ON. Whereas, at chemical potential M, the formation energy of the Wvac defect significantly increased. These results indicated that the oxygen-rich environment could promote the formation of W vacancies. Moreover, under these conditions (chemical potentials C, E, and M), the formation energy of Op, and Oc is very high, > 9 eV, which means that it is difficult to dope C and P into Bi2WO6. For PBE results, as shown in Figure S2 of the Supporting Information, the transition trends of formation energy as a function of the Fermi level are almost identical with the transition trends calculated by the HSE method. However, the energy of the pinned Fermi level is different. This is because the PBE method underestimates the band gap of Bi2WO6.

However, in Figure 1b,d, under the oxygen-poor/bismuth-rich conditions (chemical potentials D and F), we find that the formation energy of Bivac increases evidently when the formation energy of ON and OS both decrease. Thus, the lowest-formation energy defects are VO as EF < 1.55 eV. When EF > 1.55 eV formation energies of Ovac and OS are almost identical, which is the lowest-formation energy defects between these ranges. We notice that the formation energy of ON is also low in the same range of EF. Therefore, it is easy to synthesize N-doped bismuth tungstate experimentally.30,31,41 Under these conditions, the Fermi level is pinned by Ovac and ON, which is to be at ∼0.79 eV below CBM, within the n-type region. Meanwhile, the formation energy of the Wvac defect is the highest than others for the Fermi level within the band gap. Moreover, OP’s and OC’s formation energies both decrease, but the value is about 6 eV, which is still very high. These results suggested that Wvac, OP, and OC defects are unlikely to form.

The transition (ionization) levels of low-energy defects depend on free carrier generation and its contribution to the electrical conductivity,42 which can be derived according to the turning points in Figure 1. We are concerned about the transition level of Bivac, Ovac, and ON defects due to lower formation energy compared with the other defects. The ε(−/0) thermodynamic transition level (ionized from neutral to −1 charged) of Bivac is located at 0.17 eV above VBM, thereby demonstrating its shallow donor character, which agrees with the experimental observation.15 For Ovac, the donor (+2/+1) transition level is observed at 0.45 eV above VBM, while (+1/0) donor levels appear to be localized deep into the band gap 1.50 eV above VBM. Unlike the Ovac, the density of states (DOS) for Ovac2+ show delocalized distribution (we will address this issue in Section 2.2). This suggests that the Ovac acceptor state is not a deep, localized state. Such shallow levels not only extend the absorption spectra of Ovac2+ in Bi2WO6 but also play a role as trap levels, which increase the carrier lifetime.43,44 For all the chemical potentials, the formation energy of the donor Ovac is the lowest when the conductivity is p-type. Therefore, the ionization of the dominant Ovac2+ with a high concentration can produce an evident amount of hole carriers, resulting in good p-type conductivity. Moreover, the acceptor (−1/0) level associated with ON appears to be localized deep at 1.38 eV above VBM, which may act as recombination centers.

2.2. Electronic Structure with HSE06

In Figure 2, we present the total density of states (TDOS) and partial density of states (PDOS) for different native defects of Bi2WO6 between −5 and 6 eV using the HSE06 functional. We focus on the defects Ovac, Ovac+, Ovac, Bivac2–, Bivac, NO, NO, and SO. This is because these defects have the lowest formation energies under different chemical conditions for a certain range of Fermi level. The electronic properties of pure Bi2WO6 were also calculated to compare with native defects of Bi2WO6, as shown in Figure 2a. For pure Bi2WO6, the valence band maximum (VBM) of the Bi2WO6 is predominantly composed of O 2p states, which hybridize with the Bi 6s 6p and W 5d states of relatively lower density. Unlike the VBM, the conduction band minimum (CBM) has nearly identical contributions from the O 2p orbitals and the W 5d orbitals. These results are in good agreement with previous theoretical results.28,36,45

Figure 2.

Figure 2

Spin-polarized density of the states (DOS) for the (a) bulk Bi2WO6, (b) Bivac2–, (c) Bivac, (d) Ovac (e) Ovac1+, and (f) Ovac defect. The red dotted lines denote the Fermi level. Positive (negative) values of the DOS denote spin up (down).

For the Bivac2– defect, as shown in Figure 2b, we find that an occupied state in the gap (at about 0.1 eV below the Fermi level) and an isolated empty band (at about 0.9 eV above the Fermi level). These two states are mainly composed of O 2p orbitals, which may act as recombination centers when electron–hole pairs are generated. This is consistent with a previous theoretical result,15 where they only focus on the neutral bismuth vacancy. Moreover, there is a shallow defect level underneath the CBM concerning that of the pure Bi2WO6. An attempt to add an electron, for the Bivac defect, we observe from Figure 2c that the Fermi level is at the edge of the conduction band, which exhibits typical n-type characteristics. There are two occupied states (one is spin up, and another is spin down) near the Fermi level. Its localization character is fragile. This suggests that the band gap of Bivac3–:Bi2WO6 is narrower than the band gap of undoped Bi2WO6, and thus the photoabsorption probability increases.

In addition, these delocalization characters of the defect level and enhanced DOS will favor the electron transition into the conduction band under the irradiation, leading to a higher carrier concentration and an increasing electronic conductivity.46,47 It has been reported that bismuth vacancy-rich Bi2WO6 exhibits significantly increased visible-light photocatalytic oxygen evolution activity than that of pristine Bi2WO6.15 This activity can be explained by the formation of the Bivac3– defect. This is because not only the formation energy of Bivac is the lowest under oxygen-rich conditions when EF > 2.31 eV, but also it has a narrow band gap, such that its absorption is increased. More importantly, there is no recombination centers that occurred.

For the neutral oxygen vacancy, as shown in Figure 2d, there are two spin-degenerated occupied states (the positive is spin up, and the negative is spin down) in the vicinity of midgap. These two states are dominated by Bi 6p orbitals, serving as a recombination center, which decreases the efficiency of charge separation. As shown in Figure 2e, the DOS of Ovac1+ is similar to the DOS of Bivac. There is an occupied state in the gap (at about 0.1 eV below the Fermi level) and an isolated empty band (at about 0.9 eV above the Fermi level), which also act as recombination centers. However, as shown in Figure 2f, for Ovac2+, the Fermi level is at the edge of the valence band, which exhibits typical p-type characteristics. There are two empty states (one is spin up, and another is spin down) near the Fermi level, which shows the delocalization character. This suggests that the band gap of Ovac/Bi2WO6 is narrower than the band gap of pristine Bi2WO6, and its absorption is increased. The Ovac2– defect can enhance the photocatalytic activity, while Ovac and Ovac degrade, and it is desirable to induce the former one as much as possible and suppress the latter. This can be achieved by controlling EF < 0.45 eV under any environment (Figure 2), as discussed in the previous section.

In Figure 3, we plot the TDOS and PDOS for different nonmetal elements (N and S) doping into Bi2WO6 between −5 and 6 eV using the HSE06 functional. For the NO defect, as shown in Figure 3a, we found an isolated empty band (at about 1.0 eV above the Fermi level). This state is mainly composed of N 2p orbitals, which may act as recombination centers when electron–hole pairs are generated. This is consistent with the localized deep (−1/0) transition level of ON that appears to be at 1.38 eV above VBM. Based on the previous analysis, this defect is not natural to be formed because the formation energy of NO is higher than that of compensated defect Ovac under any conditions. When one more electron is injected, the empty band of ON is occupied by the extra electron (Figure 3b). The occupied states are moved below the Fermi level. These states are mainly composed of N 2p orbitals, which results in a higher degree of delocalization and a correspondingly reduced band gap of pure Bi2WO6. As shown in Figure 2, under the oxygen-poor/bismuth-rich conditions, the formation energy of NO is almost the same as the lowest formation energy (Ovac and OS). Therefore, experimentally synthesized N-doped Bi2WO6 samples30,31,41 may be NO. This explains the experimental observations of a better photocatalysts efficiency performance of N-doped Bi2WO6.

Figure 3.

Figure 3

Spin-polarized DOS for the (a) NO, (b) NO1–, and (c) SO defects using the HSE06 functional. The red dotted lines indicate the Fermi level. Positive (negative) values of the DOS denote spin up (down).

It is shown that the band gap of S-doped Bi2WO6 is about 2.33 eV in Figure 4c, which is reduced by 0.46 eV comparing with pure Bi2WO6. The substitution S mainly modifies the valence band edge and forms the O 2p and S 2p hybridization states on the valence band’s band edge. Thus, the S doping narrows the band gap of Bi2WO6 without localized gap states, which yields a clean band gap by maintaining the stability of the photocatalyst and avoids the formation of the recombination centers, leading to better photocatalytic performance. Therefore, the S–Bi2WO6 may be good candidates for visible-light photocatalysis.

Figure 4.

Figure 4

Square of the wave functions corresponding to the defect states near the Fermi level in the band gap created by (a) Bivac2–, (b) Bivac, (c) Ovac, (d) Ovac1+, (e) Ovac, (f) NO, (g) NO1–, and (h) SO defects. The isosurface value is 0.003 e/Å3. The purple, gray, red, blue, and yellow spheres represent Bi, W, O, N, and S atoms, respectively.

It is known that the shallow levels have delocalized wave functions. On the other hand, deep levels have localized ones.4850 As shown in Figure 4, we have plotted the wave-function squares of the Bivac2–, Bivac, Ovac, Ovac1+, Ovac, NO, NO1–, and SO defect states near the Fermi level in the band gap. Figure 4b,e,h plots the wave-function square of the Bivac, Ovac2+, and SO defect states. The wave function of the defect states distributes not only around the defects but also around O atoms away from the W atom, indicating a delocalized feature, which is consistent with the result that Bivac, Ovac2+, and SO are shallow defects near the Fermi level in the band gap. Figure 4a,c,d,f,g plots the wave-function square of the Bivac, Ovac, Ovac1+, NO, and NO defect states near the Fermi level in the band gap, respectively. It is seen that the wave function is localized around defects, consistent with the deep-level feature.

Generally, the conduction band (CB) and valence band (VB) edge potentials of a semiconductor play a vital role in the photocatalysis process. The Mulliken electronegativity theory51 can predict the conduction band and valence band potentials of Bi2WO6: ECB = χ – Ec – 0.5Eg, (or EVB = χ – Ec + 0.5Eg), where ECB (EVB) is the conduction (valence) band potential, the χ is the absolute electronegativity of bulk Bi2WO6, Ec is the energy of the free electron in the hydrogen scale (approximately 4.5 eV),52 and Eg is the band gap energy of the Bi2WO6 (2.79 eV). Using the Mulliken electronegativities, the band position and photoelectric thresholds for many compounds have been calculated.5355

For the Mulliken electronegativity (χ) of compound AaBbCc can be calculated according to the following equation:56,57Inline graphic, where χ (A), χ (B), and χ (C) are the absolute electronegativity of the A atoms, B atoms, and C atoms respectively; the a, b, c are the number of A atoms, B atoms, C atoms in an AaBbCc compound, respectively. According to the Mulliken definition, per atom’s absolute electronegativity is equal to the arithmetic mean of the atomic electron affinity (A) and the first ionization energy (I).56 From these data, we obtained the Mulliken electronegativity of Bi, W, O, N, and S that are 4.12, 4.40, 7.54, 7.24, and 6.22 eV, respectively.58,59 The χ value for Bi2WO6 is 6.2 eV. Hence, the ECB value of Bi2WO6 was calculated to be +0.31 eV, and the EVB value was estimated to be +3.10 eV, which agreed well with the previous calculation.60

To evaluate the defects’ influence on the photocatalytic activity of Bi2WO6, the CBM and VBM of the Bivac3–, Ovac, and SO defects concerning the Bi2WO6 are presented in Figure 5. As sown in Figure 5, for the Bivac3– defect, the VBM is raised by 0.08 eV, and the CBM is lowered by 0.03 eV relative to that of the pure Bi2WO6. This result indicated that the oxidizing capacity of the VB and the reducing capacity of the CB are reduced. Meanwhile, some defect states are introduced in the band gap, and the electron transition among the VB, CB, and these gap states leads to the visible-light absorption. For Ovac defects, the CBM does not change, and the VBM is reduced by 0.24 eV relative to that of the pure Bi2WO6, which suggests that the oxidizing capacity of VB will decrease. This is due to the band gap is reduced to 2.84 eV. Only one spin up and one spin down defect states appear in the band gap. Immensely, for the SO defect, the CBM does not change, and the VBM is raised by 0.25 eV relative to that of the pure Bi2WO6. This result suggested that the oxidizing capacity of the VB is improved. One occupied state and one unoccupied state are introduced in the band gap, improving the visible-light absorption capacity of Bi2WO6.

Figure 5.

Figure 5

Calculated VBM and CBM positions of pure Bi2WO6, Bivac3–, Ovac, and SO defects in Bi2WO6 obtained from HSE06 calculations. The VBM and CBM values are given concerning the normal hydrogen electrode (NHE) potential. The black and green arrowheads represent the occupied and unoccupied states of defect states, respectively.

2.3. Optical Properties

Generally, the optical absorption properties of photocatalytic semiconductor material are closely related to its electronic band structure. This is a significant factor affecting the photocatalytic activity.54 In this study, we focus on the Bivac3–, Ovac, and SO defects. This is because these defects narrow the band gap of Bi2WO6, and no recombination centers are formed. The calculated absorption coefficients along the [001] direction for Bivac3–, Ovac, and SO defects in Bi2WO6 are shown in Figure 6. For comparison, the absorption coefficient of Bi2WO6 was also calculated. Figure 6 indicates that pure Bi2WO6 cannot absorb the visible light effectively due to its wide band gap. After Bivac3–, Ovac, and SO doping, the absorption coefficients in the visible-light region are enhanced, and a redshift is observed on the absorption edge. The high optical absorption results from the narrow band gap and the weak electron–hole recombination, as mentioned already. In the energy region between 1.5 and 2.75 eV, the absorption coefficients of Bivac3–, and Ovac are more extensive than that of SO. When the energy is higher than 2.8 eV, the absorption coefficient is enhanced by about four times as compared with pure Bi2WO6. This is caused by absorption between the 2p orbital of O in the valence bands and the 5d orbital of W in the conduction bands according to the PDOS in Figures 2 and 3. Therefore, the enhanced absorption of incident photons will enhance the photocatalytic efficiency of Bivac3–, Ovac, and SO doping Bi2WO6 in the near-ultraviolet region. These results are also in good agreement with the electronic properties mentioned above and previous experimental reports.

Figure 6.

Figure 6

Calculated frequency-dependent optical absorption spectra for the pure Bi2WO6, Bivac3–, Ovac, and SO defects using the HES06 method.

3. Conclusions

Using the HSE06 hybrid approach, we calculate the formation energies, electronic structure, and optical properties of native defects and nonmetals (C, N, S, and P) doping in a Bi2WO6 photocatalyst. We find that the formation energies strongly depend on the impurity charge states and Fermi level position. The calculated formation energies and transition levels demonstrate that Bivac, Ovac, SO, and NO have relatively low formation energies. However, CO and PO are more challenging to form than SO and NO due to the higher formation energies. The calculated electronic structures of Bivac, Ovac, SO, and NO defects reveal that the Ovac2+, Bivac, and SO defects facilitate the delocalization of electrons. This leads to narrow band gaps of Bi2WO6 and hinders the formation of recombination centers, which would improve the photocatalysis performance of Bi2WO6. We reveal that the semiconductor nature of Bi2WO6 can be tuned from p-type to n-type by changing the growth conditions. The calculated absorption coefficients of Ovac2+, Bivac, and SO show strong absorption in the visible region, which is in good agreement with the experimental results. Based on these results, we suggest that Ovac2+, Bivac, and SO doping are good dopant candidates to tailor the visible-light absorption property of Bi2WO6 photocatalysts. Our results provide a comprehensive investigation of the charged defect of the Bi2WO6, which provides a useful strategy for further optimizing the photocatalytic performance of Bi2WO6 and its future applications.

4. Computational Methods

First-principles calculations based on the density functional theory (DFT) were performed using the projected augmented wave (PAW)61 method implemented in the Vienna ab initio simulation package (VASP).62,63 The Kohn–Sham one-electron states are expanded using the plane-wave basis set with a kinetic energy cutoff of 500 eV. The Perdew–Burke–Ernzerhof (PBE)64 exchange-correlation (XC) functional within the GGA was employed for the geometrical optimization. Since the GGA approach usually underestimates semiconductors’ band gap, we used the screened hybrid functional of Heyd, Scuseria, and Ernzerhof (HSE)38,39 for a more accurate description of the electronic structure. We tested the energy gap of pristine Bi2WO6 by adjusting the Hartree–Fock mixing parameter from 0.15 to 0.35. It was found that when α = 0.15, the calculated band gap is 2.79 eV for Bi2WO6, closely to the experimental band gap 2.80 eV.13 Therefore, in our calculations, we chose the mixing parameter equals 0.15 for the short-range Hartree–Fock exchange instead of the commonly used value of 0.25. The PAW potentials with the valence electrons 6s26p3 for Bi, 5d46s2 for W, 2s22p4 for O, 2s22p2 for C, 2s22p3 for N, 3s23p3 for P, and 3s23p4 for S have been employed. The Monkhorst–Pack k-point meshes65 for the 36-atom pure Bi2WO6 primitive cell is 6 × 2 × 6. Vacancy and doping defects calculations were performed in a 2 × 1 × 2 supercell (144 atoms). For these systems, we used a 2 × 1 × 2 k-point, which was found to be sufficient to reach convergence for bulk calculations and used for geometry optimization and electronic property calculations. For defects calculations, the supercells’ sizes were kept, but the atomic positions were relaxed until forces were smaller than 0.03 eV/Å, corresponding to the energy convergence smaller than 0.0001 eV. For all calculations, spin polarization was taken into consideration. The experimental atomic positions66 were employed as the starting points of relaxation, and the following calculations were performed using the relaxed atomic positions.

The formation energy of a defect X in charge state q is defined as67,68

4. 1

where Etot(Xq) is the total energy of the Bi2WO6 in the presence of a defect X in charge state q in the supercell, and Etot(bulk) is the total energy of pristine Bi2WO6 in the absence of a defect in the supercell. For charged defects (q ≠ 0), the formation energy depends on the Fermi level (EF), which varies from the valence-band maximum (VBM) to the calculated conduction-band minimum (CBM). In this study, EF at the VBM is set to 0 eV, and CBM is set to 2.79 eV, respectively. EV is the energy cost in removing an electron from the pure supercell. ΔV is determined by aligning O atoms’ 1s core levels far away from the defect center in defective cells to the pure bulk. Additionally, Ecorr is a correction term to account for supercell errors such as image–charge interactions. This is because of the supercell and periodical boundary conditions used in the DFT calculations. Following the approach given by Lany and Zunger,69,70 the image–charge correction term is written as

4. 2

where f is a shape factor related to the supercell geometries,70q is the total charge, α is the Madelung constant, ε is the static dielectric constant of the pristine bulk,71 and L is the linear dimension of the supercell (L = V1/3, V is the volume of the supercell). ni corresponds to the number of atoms added to (ni > 0) or taken from (ni < 0) the system to form the X. Defects with lower formation energies will form more easily and occur in higher concentrations. μi = Ei + Δμi is the chemical potentials of the constituent i in the reservoir, where Ei denotes the calculated standard state energies.72 The values of Δμi can vary depending on the environmental conditions in thermodynamic equilibrium but are restricted by the relation12

4. 3

where ΔHf(Bi2WO6) is the formation enthalpy of per formula of Bi2WO6, which is equal to −14.98 eV according to our HSE06 calculations. This result is very close to previous −15.50 eV estimates using the PBE method12 but higher than the experimental value of −11.52 eV.73 The upper bound is Δμi ≤ 0 as, at this point, precipitation of element to its standard state occurs. Thereby, the non-formation of the pure solids Bi and W or O2 gas, requires

4. 4

The following relationships should be satisfied to avoid the formation of other oxide phases (such as WO2, WO3, and Bi2O3):

4. 5
4. 6
4. 7

Eqs 37 determine the phase diagram of Bi2WO6 as a two-dimensional panel with two independent variables, ΔμBi, and ΔμW. The PBE- and HSE06-calculated stable regions for Bi2WO6 in the phase diagram are shown as the light green and light blue shaded quadrilateral in Figure S1 and Figure 7, respectively. The values inside of these regions induce no precipitation of the second phase. Our PBE result is in excellent agreement with the previous theoretical studies.12,74,75 The stability regions are estimated by the PBE and HSE06 methods, having almost the same shape and size. Thus, the coordinates of C, D, E, and F in Figure S1 are not much different from those of C, D, E, and F in Figure 7. This indicates that the calculated results of the two methods agree with each other.

Figure 7.

Figure 7

Ranges of chemical potentials showing the formation of different competing phases under thermodynamic equilibrium calculated by HSE06. The light-blue-shaded quadrilateral CDEF is the region where Bi2WO6 is stable without forming other phases. Specific points C, D, E, F, and M are chosen as the representative chemical potentials for the following defect formation energy calculation.

Five representative chemical potentials points C, D, E, F, and M were selected in the light-blue-shaded regions to calculate defect formation energies. C and E represent oxygen-rich conditions, D and F indicate oxygen-poor and bismuth-rich conditions, and M in the stable region represents condition between O-rich conditions and O-poor/bi-rich conditions, respectively. Moreover, we obtained the values of ΔμBi, ΔμW, and ΔμO and then calculated the values of corresponding chemical potentials using μi = Ei + Δμi; the results are listed in Table 2.

Table 2. Chemical Potentials Used to Represent Growth Conditions (C, D, E, F, and M ) for Bi2WO6a.

  chemical potentials (eV)
condition μBi μW μO μC μN μP μS
C –7.62 –22.63 –6.16 –13.87 –9.68 –13.21 –8.88
D –4.29 –15.97 –8.38 –9.96 –9.47 –7.66 –5.34
E –7.22 –23.43 –6.16 –13.87 –9.68 –13.21 –8.88
F –4.29 –17.59 –8.11 –9.97 –9.47 –8.34 –5.34
M –6.29 –20.31 –6.99 –12.21 –9.47 –11.14 –6.39
a

M is one point in the Bi2WO6 stability region.

When considering impurity (C, N, P, and S) doping, impurities’ chemical potentials should also be considered. The following constraints are enforced to avoid the formation of impurity-related phases:

4. 8
4. 9
4. 10
4. 11
4. 12
4. 13

For example, for N doping at C chemical potentials, ΔμN= −0.21 eV. In this case, the μN= −9.68 eV is used for the calculation of formation energy for N-doping defects. The derived upper-limit chemical potentials for C, N, P, and S by HSE06 are listed in Table 1, considering the special growth conditions.

Table 1. Calculated Enthalpy of Formation Per Formula Unit of Bi2WO6 and the Possible Limiting Phases by HSE06 Calculationa.

compound ΔHf (eV) experiment
Bi2WO6 –14.98 –11.5273
Bi2O3 –5.86 –5.9576
WO2 –5.75 –6.1176
WO3 –8.32 –8.7476
CO –0.71 –1.1476
CO2 –3.91 –4.0876
P2O5 –14.38  
SO2 –3.06 –3.0876
SO3 –4.13  
Bi2S3 –1.78 –1.4876
WC –0.41  
N2O5 –0.42 –0.4576
a

Experimental values at 298 K are provided for comparison (unit: eV).

The formation energy (given per formula unit) calculated by HSE06 for the various possible competing phases (Bi2O3, WO2, WO3, CO, CO2, P2O5, SO2, SO3, and Bi2S3 WC N2O5) are shown in Table1 together with the available experimental values. It is seen that the calculated values are in good agreement with the experimental results.

We further calculated the thermodynamic transition energy levels ε(q/q ′ ) of the defect in Bi2WO6 using the following expression:

4. 14

where Ef(Xq) and Ef(Xq) represent formation energies of the doped Bi2WO6 in the charge states q and q′, respectively. Ε(q/q′) refers to the Fermi level at which the formation energies of the defect in charge q and q′ cross.

The imaginary part ε2(ω) of the dielectric function ε(ω) is calculated using the standard formulation

4. 15

where V is the cell volume, ω is the energy of the incident photon, p→ is the momentum operator, |nk→> denotes the electronic state k→ in band n, and fnk→ is the Fermi occupation function. The real part ε1(ω) is related to ε2(ω) by the Kramer–Krönig transformation. The absorption coefficient α(ω) can be derived from ε1(ω) and ε2(ω) as follows:77

4. 16

It should be noted that many-body perturbation GW (“G” is the one-body Green’s function, which describes the propagation of a particle in an interacting system, and “W” is the linear response dynamically screened Coulomb interaction) approximation is used to calculate quasiparticle excitations in semiconductors and insulators, as measured by direct and inverse photoemission experiments.78 For the optical spectrum, the Bethe–Salpeter equation (BSE) is solved to include the effects of electron–hole interactions.79 However, more GW and BSE calculations for doped Bi2WO6 are limited by their expensive computational time in this paper.

Acknowledgments

This research was supported by the Natural Science Foundation of Guizhou Province (grant no. QHPT [2017]5790-02), by the National Natural Science Foundation of China (grant no. 11804078), and support from Henan university (grant no. CJ3050A0240050). The Supercomputing Center of the University of Science and Technology of China is also acknowledged.

Supporting Information Available

The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acsomega.0c03685.

  • Ranges of chemical potentials for Bi2WO6 under thermodynamic equilibrium calculated by PBE and defect formation energy calculations of Bi2WO6 supercells with various types of point defects in different charge states using the PBE-DFT method as a function of the Fermi level (PDF)

Author Contributions

§ J.Z. and P.D. contributed equally to this paper.

The authors declare no competing financial interest.

Supplementary Material

ao0c03685_si_001.pdf (96.3KB, pdf)

References

  1. Ran J.; Zhang J.; Yu J.; Jaroniec M.; Qiao S. Z. Earth-abundant cocatalysts for semiconductor-based photocatalytic water splitting. Chem. Soc. Rev. 2014, 43, 7787–7812. 10.1039/C3CS60425J. [DOI] [PubMed] [Google Scholar]
  2. Chen X.; Shen S.; Guo L.; Mao S. S. Semiconductor-based Photocatalytic Hydrogen Generation. Chem. Rev. 2010, 110, 6503–6570. 10.1021/cr1001645. [DOI] [PubMed] [Google Scholar]
  3. Chen C.; Ma W.; Zhao J. Semiconductor-mediated photodegradation of pollutants under visible-light irradiation. Chem. Soc. Rev. 2010, 39, 4206–4219. 10.1039/b921692h. [DOI] [PubMed] [Google Scholar]
  4. Hisatomi T.; Kubota J.; Domen K. Recent advances in semiconductors for photocatalytic and photoelectrochemical water splitting. Chem. Soc. Rev. 2014, 43, 7520–7535. 10.1039/C3CS60378D. [DOI] [PubMed] [Google Scholar]
  5. Pan H. Principles on design and fabrication of nanomaterials as photocatalysts for water-splitting. Renewable Sustainable Energy Rev. 2016, 57, 584–601. 10.1016/j.rser.2015.12.117. [DOI] [Google Scholar]
  6. Jiang J.; Zhao K.; Xiao X.; Zhang L. Synthesis and Facet-dependent Photoreactivity of BiOCl Single-crystalline Nanosheets. J. Am. Chem. Soc. 2012, 134, 4473–4476. 10.1021/ja210484t. [DOI] [PubMed] [Google Scholar]
  7. Cheng H.; Huang B.; Dai Y. Engineering BiOX (X = Cl, Br, I) Nanostructures for Highly Efficient Photocatalytic Applications. Nanoscale 2014, 6, 2009–2026. 10.1039/c3nr05529a. [DOI] [PubMed] [Google Scholar]
  8. Di J.; Xia J.; Li H.; Guo S.; Dai S. Bismuth oxyhalide layered materials for energy and environmental applications. Nano Energy 2017, 41, 172–192. 10.1016/j.nanoen.2017.09.008. [DOI] [Google Scholar]
  9. Yao W.; Zhang J.; Wang Y.; Ren F. Hybrid Density Functional Study on the Mechanism for the Enhanced Photocatalytic Properties of the Ultrathin Hybrid Layered Nanocomposite g-C3N4/BiOCl. Appl. Surf. Sci. 2018, 435, 1351–1360. 10.1016/j.apsusc.2017.11.259. [DOI] [Google Scholar]
  10. Liu X.; Gu S.; Zhao Y.; Zhou G.; Li W. BiVO4, Bi2WO6 and Bi2MoO6 photocatalysis: A brief review. J. Mater. Sci. Technol. 2020, 56, 45–68. 10.1016/j.jmst.2020.04.023. [DOI] [Google Scholar]
  11. Zhang N.; Ciriminna R.; Pagliaro M.; Xu Y.-J. Nanochemistry-derived Bi2WO6 Nanostructures: Towards Production of Sustainable Chemicals and Fuels Induced by Visible Light. Chem. Soc. Rev. 2014, 43, 5276–5287. 10.1039/C4CS00056K. [DOI] [PubMed] [Google Scholar]
  12. Ren F.; Zhang J.; Wang Y.; Yao W. A Graphene-coupled Bi2WO6 Nanocomposite with Enhanced Photocatalytic Performance: a First-principles Study. Phys. Chem. Chem. Phys. 2016, 18, 14113–14121. 10.1039/C6CP00458J. [DOI] [PubMed] [Google Scholar]
  13. Kudo A.; Hijii S. H2 or O2 Evolution from Aqueous Solutions on Layered Oxide Photocatalysts Consisting of Bi3+ with 6s2 Configuration and d0 Transition Metal Ions. Chem. Lett. 1999, 28, 1103–1104. 10.1246/cl.1999.1103. [DOI] [Google Scholar]
  14. Li C.; Chen G.; Sun J.; Rao J.; Han Z.; Hu Y.; Zhou Y. A Novel Mesoporous Single-Crystal-Like Bi2WO6 with Enhanced Photocatalytic Activity for Pollutants Degradation and Oxygen Production. ACS Appl. Mater. Interfaces 2015, 7, 25716–25724. 10.1021/acsami.5b06995. [DOI] [PubMed] [Google Scholar]
  15. Di J.; Chen C.; Zhu C.; Ji M.; Xia J.; Yan C.; Hao W.; Li S.; Li H.; Liu Z. Bismuth vacancy mediated single unit cell Bi2WO6 nanosheets for boosting photocatalytic oxygen evolution. Appl. Catal. B 2018, 238, 119–125. 10.1016/j.apcatb.2018.06.066. [DOI] [Google Scholar]
  16. Kang D.; Park Y.; Hill J. C.; Choi K. S. Preparation of Bi-Based Ternary Oxide Photoanodes BiVO4, Bi2WO6, and Bi2Mo3O12 Using Dendritic Bi Metal Electrodes. J. Phys. Chem. Lett. 2014, 5, 2994–2999. 10.1021/jz501544k. [DOI] [PubMed] [Google Scholar]
  17. Zhou Y.; Zhang Y.; Lin M.; Long J.; Zhang Z.; Lin H.; Wu J. C.-S.; Wang X. Monolayered Bi2WO6 nanosheets mimicking heterojunction interface with open surfaces for photocatalysis. Nat. Commun. 2015, 6, 8340. 10.1038/ncomms9340. [DOI] [PMC free article] [PubMed] [Google Scholar]
  18. Zhang L.; Zhu Y. A review of controllable synthesis and enhancement of performances of bismuth tungstate visible-light-driven photocatalysts. Catal. Sci. Technol. 2012, 2, 694. 10.1039/c2cy00411a. [DOI] [Google Scholar]
  19. Zhang L.; Wang H.; Chen Z.; Wong P. K.; Liu J. Bi2WO6 micro/nano-structures: Synthesis, modifications and visible-light-driven photocatalytic applications. Appl. Catal. B 2011, 106, 1–13. [Google Scholar]
  20. Tian J.; Sang Y.; Yu G.; Jiang H.; Mu X.; Liu H. A Bi2WO6-based hybrid photocatalyst with broad spectrum photocatalytic properties under UV, visible, and near-infrared irradiation. Adv. Mater. 2013, 25, 5075–5080. 10.1002/adma.201302014. [DOI] [PubMed] [Google Scholar]
  21. Huang H.; Cao R.; Yu S.; Xu K.; Hao W.; Wang Y.; Dong F.; Zhang T.; Zhang Y. Single-unit-cell layer established Bi2WO6 3D hierarchical architectures: Efficient adsorption, photocatalysis and dye-sensitized photoelectrochemical performance. Appl. Catal. B 2017, 219, 526–537. 10.1016/j.apcatb.2017.07.084. [DOI] [Google Scholar]
  22. Shang Y.; Cui Y.; Shi R.; Yang P. Effect of acetic acid on morphology of Bi2WO6 with enhanced photocatalytic activity. Mater. Sci. Semicond. Process. 2019, 89, 240–249. 10.1016/j.mssp.2018.09.026. [DOI] [Google Scholar]
  23. Meng J.; Xiong X.; Zhang X.; Xu Y. Improved photocatalytic degradation of chlorophenol over Pt/Bi2WO6 on addition of phosphate. Appl. Surf. Sci. 2018, 439, 859–867. 10.1016/j.apsusc.2017.12.247. [DOI] [Google Scholar]
  24. Zhang L.; Man Y.; Zhu Y. Effects of Mo Replacement on the Structure and Visible-Light-Induced Photocatalytic Performances of Bi2WO6 Photocatalyst. ACS Catal. 2011, 1, 841–848. 10.1021/cs200155z. [DOI] [Google Scholar]
  25. Li H.; Hao H.; Jin S.; Guo W.; Hu X.; Hou H.; Zhang G.; Yan S.; Gao W.; Liu G. Synthesis of Yb3+ /Ho3+ co-doped Bi2WO6 upconversion photocatalyst with highly improved visible light photocatalytic activity. Catal. Commun. 2017, 97, 60–64. 10.1016/j.catcom.2017.04.021. [DOI] [Google Scholar]
  26. Marschall R.; Wang L. Non-metal doping of transition metal oxides for visible-light photocatalysis. Catal. Today 2014, 225, 111–135. 10.1016/j.cattod.2013.10.088. [DOI] [Google Scholar]
  27. Hu T.; Li H.; Du N.; Hou W. Iron-Doped Bismuth Tungstate with an Excellent Photocatalytic Performance. ChemCatChem 2018, 10, 3040–3048. 10.1002/cctc.201701965. [DOI] [Google Scholar]
  28. Lai K.; Zhu Y.; Lu J.; Dai Y.; Huang B. N- and Mo-doping Bi2WO6 in photocatalytic water splitting. Comput. Mater. Sci. 2013, 67, 88–92. 10.1016/j.commatsci.2012.08.024. [DOI] [Google Scholar]
  29. Fu Y.; Chang C.; Chen P.; Chu X.; Zhu L. Enhanced photocatalytic performance of boron doped Bi2WO6 nanosheets under simulated solar light irradiation. J. Hazard. Mater. 2013, 254-255, 185–192. 10.1016/j.jhazmat.2013.03.046. [DOI] [PubMed] [Google Scholar]
  30. Hoang L. H.; Phu N. D.; Peng H.; Chen X.-B. High photocatalytic activity N-doped Bi2WO6 nanoparticles using a two-step microwave-assisted and hydrothermal synthesis. J. Alloys Compd. 2018, 744, 228–233. 10.1016/j.jallcom.2018.02.094. [DOI] [Google Scholar]
  31. Shang M.; Wang W.; Zhang L.; Xu H. Bi2WO6 with significantly enhanced photocatalytic activities by nitrogen doping. Mater. Chem. Phys. 2010, 120, 155–159. 10.1016/j.matchemphys.2009.10.039. [DOI] [Google Scholar]
  32. Dumrongrojthanath P.; Phuruangrat A.; Thongtem S.; Thongtem T. Hydrothermal preparation of visible-light-driven Br-doped Bi2WO6 photocatalyst. Mater. Lett. 2017, 209, 501–504. 10.1016/j.matlet.2017.08.089. [DOI] [Google Scholar]
  33. Kumar B. V.; Prasad M. D.; Vithal M. Enhanced visible light photocatalytic activity of Sn doped Bi2WO6 nanocrystals. Mater. Lett. 2015, 152, 200–202. 10.1016/j.matlet.2015.03.096. [DOI] [Google Scholar]
  34. Huo W. C.; Dong X.’a.; Li J. Y.; Liu M.; Liu X. Y.; Zhang Y. X.; Dong F. Synthesis of Bi2WO6 with gradient oxygen vacancies for highly photocatalytic NO oxidation and mechanism study. Chem. Eng. J. 2019, 361, 129–138. 10.1016/j.cej.2018.12.071. [DOI] [Google Scholar]
  35. Jing T.; Dai Y.; Wei W.; Ma X.; Huang B. Near-infrared photocatalytic activity induced by intrinsic defects in Bi2MO6 (M = W, Mo). Phys. Chem. Chem. Phys. 2014, 16, 18596–18604. 10.1039/C4CP01846J. [DOI] [PubMed] [Google Scholar]
  36. Lai K.; Wei W.; Zhu Y.; Guo M.; Dai Y.; Huang B. Effects of oxygen vacancy and N-doping on the electronic and photocatalytic properties of Bi2MO6 (M=Mo, W). J. Solid State Chem. 2012, 187, 103–108. 10.1016/j.jssc.2012.01.004. [DOI] [Google Scholar]
  37. Matsubara M.; Saniz R.; Partoens B.; Lamoen D. Doping anatase TiO2 with group V-b and VI-b transition metal atoms: a hybrid functional first-principles study. Phys. Chem. Chem. Phys. 2017, 19, 1945–1952. 10.1039/C6CP06882K. [DOI] [PubMed] [Google Scholar]
  38. Heyd J.; Scuseria G. E.; Ernzerhof M. Hybrid unctionals Based on a Screened Coulomb Potential. J. Chem. Phys. 2003, 118, 8207–8215. 10.1063/1.1564060. [DOI] [Google Scholar]
  39. Heyd J.; Scuseria G. E.; Ernzerhof M. Erratum: “Hybrid Functionals Based on a Screened Coulomb Potential” [J. Chem. Phys. 118, 8207 (2003)]. J. Chem. Phys. 2006, 124, 219906. 10.1063/1.2204597. [DOI] [Google Scholar]
  40. Freysoldt C.; Neugebauer J.; Van de Walle C. G. Electrostatic interactions between charged defects in supercells. Phys. Stat. Sol. B 2011, 248, 1067–1076. 10.1002/pssb.201046289. [DOI] [Google Scholar]
  41. Tang B.; Jiang G.; Wei Z.; Li X.; Wang X.; Jiang T.; Chen W.; Wan J. Preparation of N-Doped Bi2WO6 Microspheres for Efficient Visible Light-Induced Photocatalysis. Acta Metall. Sin. 2014, 27, 124–130. 10.1007/s40195-013-0009-z. [DOI] [Google Scholar]
  42. Chen S.; Walsh A.; Gong X.-G.; Wei S.-H. Classification of Lattice Defects in the Kesterite Cu2ZnSnS4 and Cu2ZnSnSe4 Earth-Abundant Solar Cell Absorbers. Adv. Mater. 2013, 25, 1522–1539. 10.1002/adma.201203146. [DOI] [PubMed] [Google Scholar]
  43. Brancho J. J.; Bartlett B. M. Challenges in Co-Alloyed Titanium Oxynitrides, a Promising Class of Photochemically Active Materials. Chem. Mater. 2015, 27, 7207–7217. 10.1021/acs.chemmater.5b02357. [DOI] [Google Scholar]
  44. Choi H.; Shin D.; Yeo B. C.; Song T.; Han S. S.; Park N.; Kim S. Simultaneously Controllable Doping Sites and the Activity of a W–N Codoped TiO2 Photocatalyst. ACS Catal. 2016, 6, 2745–2753. 10.1021/acscatal.6b00104. [DOI] [Google Scholar]
  45. Liu X.; Fan H.-Q. Theoretical studies on electronic structure and optical properties of Bi2WO6. Optik 2018, 158, 962–969. 10.1016/j.ijleo.2017.12.124. [DOI] [Google Scholar]
  46. Gao S.; Gu B.; Jiao X.; Sun Y.; Zu X.; Yang F.; Zhu W.; Wang C.; Feng Z.; Ye B.; Xie Y. Highly Efficient and Exceptionally Durable CO2 Photoreduction to Methanol over Freestanding Defective Single-Unit-Cell Bismuth Vanadate Layers. J. Am. Chem. Soc. 2017, 139, 3438–3445. 10.1021/jacs.6b11263. [DOI] [PubMed] [Google Scholar]
  47. Lei F.; Sun Y.; Liu K.; Gao S.; Liang L.; Pan B.; Xie Y. Oxygen vacancies confined in ultrathin indium oxide porous sheets for promoted visible-light water splitting. J. Am. Chem. Soc. 2014, 136, 6826–6829. 10.1021/ja501866r. [DOI] [PubMed] [Google Scholar]
  48. Segev D.; Wei S. H. Design of shallow donor levels in diamond by isovalent-donor coupling. Phys. Rev. Lett. 2003, 91, 126406. 10.1103/PhysRevLett.91.126406. [DOI] [PubMed] [Google Scholar]
  49. Yin W.-J.; Wei S.-H.; Al-Jassim M. M.; Turner J.; Yan Y. Doping properties of monoclinic BiVO4 studied by first-principles density-functional theory. Phys. Rev. B 2011, 83, 155102. 10.1103/PhysRevB.83.155102. [DOI] [Google Scholar]
  50. Feng C.; Yin W.-J.; Nie J.; Zu X.; Huda M. N.; Wei S.-H.; Al-Jassim M. M.; Turner J. A.; Yan Y. Strong asymmetrical doping properties of spinel CoAl2O4. J. Appl. Phys. 2012, 111, 093723 10.1063/1.4716025. [DOI] [Google Scholar]
  51. Butler M. A. Prediction of Flatband Potentials at Semiconductor-Electrolyte Interfaces from Atomic Electronegativities. J. Electrochem. Soc. 1978, 125, 228. 10.1149/1.2131419. [DOI] [Google Scholar]
  52. Morrison S. R.Electrochemistry at Semiconductor and Oxidized Metal Electrodes; Plenum Press: New York, NY, USA, 1980. [Google Scholar]
  53. Lim Y. K.; Keong Koh E. W.; Zhang Y.-W.; Pan H. Ab initio design of GaN-based photocatalyst: ZnO-codoped GaN nanotubes. J. Power Sources 2013, 232, 323–331. 10.1016/j.jpowsour.2013.01.066. [DOI] [Google Scholar]
  54. Zhang J.; Ren F.; Deng M.; Wang Y. Enhanced visible-light photocatalytic activity of a g-C3N4/BiVO4 nanocomposite: a first-principles study. Phys. Chem. Chem. Phys. 2015, 17, 10218–10226. 10.1039/C4CP06089J. [DOI] [PubMed] [Google Scholar]
  55. Zhang J.; Deng M.; Ren F.; Wu Y.; Wang Y. Effects of Mo/W Codoping on the Visible-light Photocatalytic Activity of Monoclinic BiVO4 within the GGA + U Framework. RSC Adv. 2016, 6, 12290–12297. 10.1039/C5RA22659G. [DOI] [Google Scholar]
  56. El Halouani F.; Deschanvres A. Interfaces semi-conducteur-electrolyte: Correlations entre le potentiel de bande plate et les echelles d’electronegativite. Mater. Res. Bull. 1982, 17, 1045–1052. 10.1016/0025-5408(82)90130-1. [DOI] [Google Scholar]
  57. Gao J.; Zeng W.; Tang B.; Zhong M.; Liu Q.-J. Density functional characterization of Bi-based photocatalysts: BiTaO4, Bi4Ta2O11 and Bi7Ta3O18. Mater. Sci. Semicond. Process. 2021, 121, 105447. 10.1016/j.mssp.2020.105447. [DOI] [Google Scholar]
  58. Andersen T.; Haugen H. K.; Hotop H. Binding Energies in Atomic Negative Ions: III. J. Phys. Chem. Ref. Data 1999, 28, 1511–1533. 10.1063/1.556047. [DOI] [Google Scholar]
  59. Kittel C.Introduction To Solid State Physics; 8Th Edition Wiley India Pvt. Limited: 2005. [Google Scholar]
  60. Wang B.; Yang H.; Xian T.; Di L. J.; Li R. S.; Wang X. X. Synthesis of Spherical Bi2WO6 Nanoparticles by a Hydrothermal Route and Their Photocatalytic Properties. J. Nanomater. 2015, 2015, 1–7. [Google Scholar]
  61. Kresse G.; Joubert D. From ultrasoft pseudopotentials to the projector augmented-wave method. Phys. Rev. B 1999, 59, 1758. 10.1103/PhysRevB.59.1758. [DOI] [Google Scholar]
  62. Kresse G.; Hafner J. Ab initio molecular dynamics for liquid metals. Phys. Rev. B 1993, 47, 558. 10.1103/PhysRevB.47.558. [DOI] [PubMed] [Google Scholar]
  63. Kresse G.; Furthmüller J. Efficient Iterative Schemes for ab initio Total-energy Calculations using a Plane-wave Basis Set. Phys. Rev. B 1996, 54, 11169. 10.1103/PhysRevB.54.11169. [DOI] [PubMed] [Google Scholar]
  64. Perdew J. P.; Burke K.; Ernzerhof M. Generalized gradient approximation made simple. Phys. Rev. Lett. 1996, 77, 3865. 10.1103/PhysRevLett.77.3865. [DOI] [PubMed] [Google Scholar]
  65. Monkhorst H. J.; Pack J. D. Special points for brillouin-zone integrations. Phys. Rev. B 1976, 13, 5188. 10.1103/PhysRevB.13.5188. [DOI] [Google Scholar]
  66. Knight K. S. The crystal structure of russellite; a re-determination using neutron powder diffraction of synthetic Bi2WO6. Mineral. Mag. 1999, 56, 399–409. 10.1180/minmag.1992.056.384.13. [DOI] [Google Scholar]
  67. Freysoldt C.; Grabowski B.; Hickel T.; Neugebauer J.; Kresse G.; Janotti A.; Van de Walle C. G. First-principles calculations for point defects in solids. Rev. Mod. Phy. 2014, 86, 253–305. 10.1103/RevModPhys.86.253. [DOI] [Google Scholar]
  68. Zhang S. B.; Northrup J. E. Chemical potential dependence of defect formation energies in GaAs: Application to Ga self-diffusion. Phys. Rev. Lett. 1991, 67, 2339–2342. 10.1103/PhysRevLett.67.2339. [DOI] [PubMed] [Google Scholar]
  69. Lany S.; Zunger A. Assessment of correction methods for the band-gap problem and for finite-size effects in supercell defect calculations: Case studies for ZnO and GaAs. Phys. Rev. B 2008, 78, 235104. 10.1103/PhysRevB.78.235104. [DOI] [Google Scholar]
  70. Lany S.; Zunger A. Accurate prediction of defect properties in density functional supercell calculations. Modell. Simul. Mater. Sci. Eng. 2009, 17, 084002 10.1088/0965-0393/17/8/084002. [DOI] [Google Scholar]
  71. Komsa H.-P.; Rantala T. T.; Pasquarello A. Finite-size supercell correction schemes for charged defect calculations. Phys. Rev. B 2012, 86, 045112 10.1103/PhysRevB.86.045112. [DOI] [PubMed] [Google Scholar]
  72. Buckeridge J.; Scanlon D. O.; Walsh A.; Catlow C. R. A. Automated procedure to determine the thermodynamic stability of a material and the range of chemical potentials necessary for its formation relative to competing phases and compounds. Comput. Phys. Commun. 2014, 185, 330–338. 10.1016/j.cpc.2013.08.026. [DOI] [Google Scholar]
  73. Phapale S.; Das D.; Mishra R. Standard molar enthalpy of formation of Bi2WO6(s) and Bi2W2O9(s) compounds. J. Chem. Thermodyn. 2013, 63, 74–77. 10.1016/j.jct.2013.04.004. [DOI] [Google Scholar]
  74. Jing T.; Dai Y.; Ma X.; Wei W.; Huang B. The photocatalytic properties of ultrathin bismuth oxychloride nanosheets: a first principles study. Phys. Chem. Chem. Phys. 2016, 18, 7261–7268. 10.1039/C5CP07437A. [DOI] [PubMed] [Google Scholar]
  75. Morita K.; Park J.-S.; Kim S.; Yasuoka K.; Walsh A. Crystal Engineering of Bi2WO6 to Polar Aurivillius-Phase Oxyhalides. J. Phys. Chem. C 2019, 123, 29155–29161. 10.1021/acs.jpcc.9b09806. [DOI] [Google Scholar]
  76. Haynes W. M.; Lide D. R.; Bruno T. J.. CRC handbook of chemistry and physics; CRC Press: Boca Raton, Florida, 2014. [Google Scholar]
  77. Saha S.; Sinha T. P.; Mookerjee A. Electronic Structure, Chemical Bonding, and Optical Properties of Paraelectric BaTiO3. Phys. Rev. B 2000, 62, 8828. 10.1103/PhysRevB.62.8828. [DOI] [Google Scholar]
  78. Reining L. The GW approximation: content, successes and limitations. WIREs Comput. Mol. Sci. 2017, 8, e1344. [Google Scholar]
  79. Onida G.; Reining L.; Rubio A. Electronic excitations: density-functional versus many-body Green’s-function approaches. Rev. Mod. Phy. 2002, 74, 601–659. 10.1103/RevModPhys.74.601. [DOI] [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials

ao0c03685_si_001.pdf (96.3KB, pdf)

Articles from ACS Omega are provided here courtesy of American Chemical Society

RESOURCES