Kangyu
Zhang
ab,
Li-Chang
Yin
*ab,
Guoqiang
Deng
ab,
Xing-Qiu
Chen
ab,
Hui-Ming
Cheng
ac and
Gang
Liu
*ab
aShenyang National Laboratory for Materials Science, Institute of Metal Research, Chinese Academy of Sciences, 72 Wenhua Road, Shenyang 110016, China. E-mail: lcyin@imr.ac.cn; gangliu@imr.ac.cn
bSchool of Materials Science and Engineering, University of Science and Technology of China, 72 Wenhua Road, Shenyang 110016, China
cInstitute of Technology for Carbon Neutrality, Shenzhen Institute of Advanced Technology, Chinese Academy of Sciences, 1068 Xueyuan Blvd, Shenzhen 518055, China
First published on 25th September 2024
Identifying the prerequisite of inducing band-like doping levels in wide bandgap metal oxides is a crucial yet open question. Herein, taking boron (B) and nitrogen (N) codoped anatase TiO2 as an example and combining density functional theory calculation with machine learning, it has been revealed that the band-like doping levels mainly originate from a strong p–d coupling between O/N-pπ and Ti-t2g orbitals. Significantly, the existence of strong p–d coupling is an intrinsic characteristic of anatase TiO2 determined entirely by crystal symmetry, which can be used as a universal criterion to predict whether band-like doping levels can be induced in other metal oxides, and the criterion has been fully verified on rutile and brookite TiO2, ScTaO4, WO3, SnO2 and MgTa2O6. Besides, the strong p–d coupling offers a theoretical basis for the long-held empirical understanding that uniform doping is important in achieving stronger visible light absorption for wide bandgap metal oxide photocatalysts. Overall, the uncovered strong p–d coupling provides a simple yet profound guideline for bandgap engineering of metal oxides.
Nevertheless, the emergence of red TiO2 has gone beyond this traditional understanding.24,25 Specifically, by gradient boron (B) and nitrogen (N) codoping, the first red anatase TiO2 with a bandgap value of about 2 eV was obtained,24 which exhibits a band-to-band redshift of the light absorption edge compared with pristine anatase TiO2. Advancing from gradient to uniform doping,25 the photocatalytic oxygen evolution activity of red anatase TiO2 has been substantially increased, implying a much higher hole mobility contributed by band-like doping levels. Considering the fact that the atomic ratio of N to Ti is only about 5% in red anatase TiO2,25 it is much lower than the N content in oxynitrides. However, the bandgap value of anatase TiO2 can still be substantially narrowed from 3.2 to about 2 eV, which is similar to that of oxynitrides.26 This suggests the existence of certain intrinsic character allowing for band-like doping levels at a low N-doping concentration in anatase TiO2.
Elucidating this intrinsic character necessitates the identification of characteristic B/N-codoping configurations featuring band-like doping levels, which are likely to exhibit the most significant bandgap narrowing among all configurations due to the large energy level broadening of band-like doping levels. Although the bandgap can also be narrowed by localized doping levels related to direct bonding between N dopants,27 this can be easily excluded by scrutinizing the band structures. Therefore, calculating the bandgap values for all possible B/N-codoping configurations in anatase TiO2 is the precondition to efficiently and accurately identify those configurations featuring band-like doping levels. However, to reproduce the experimental doping level (N:
Ti ≈ 5 at%)24 as closely as possible and maintain the charge balance, one needs to substitute three lattice O atoms with three N atoms and introduce one additional interstitial B atom in a 3 × 3 × 1 anatase TiO2 supercell (N
:
Ti = 8.33 at%), resulting in a total of 8144 symmetrically inequivalent B/N-codoping configurations. This makes the calculation of bandgap values for all these configurations far beyond the capability of conventional density functional theory (DFT) calculations. Thus, previous studies have only been able to consider a few specific configurations based on researchers' individual intuition and/or expertise.24,28,29 Consequently, the intrinsic character of anatase TiO2 allowing band-like doping levels has remained elusive over the past two decades.
Different from conventional DFT calculations, the machine learning (ML)30–35 methods offer a promising solution to accurately and efficiently predict the bandgap values of all B/N-codoping configurations in anatase TiO2. Since applying the ML methods requires each B/N-codoping configuration to be represented by a fixed-length array, termed structure descriptor, in this work, we first proposed an accurate structure descriptor, and then trained a highly accurate ML model for predicting the bandgap values of all B/N-codoping configurations in anatase TiO2. Based on the ML predictions, B/N-codoping configurations with band-like doping levels were identified, and thorough analysis revealed that the existence of strong p–d coupling between O/N-pπ and Ti-t2g orbitals in anatase TiO2 is essential for inducing band-like doping levels. Since the existence of strong p–d coupling is entirely determined by crystal symmetry, we further generalized it as a universal criterion to predict whether band-like doping levels can be induced in other metal oxides, and the criterion has been fully verified on rutile and brookite TiO2, ScTaO4, WO3, SnO2 and MgTa2O6.
Based on the ML prediction, 20 B/N-codoping configurations with the most significant bandgap narrowing have been identified, which exhibit distinctive periodic features of dopant spatial orderings. Specifically, all three N dopants are located on the same (010) atomic plane and exhibit two characteristic spatial orderings, while the interstitial B dopant can occupy any interstitial site on either the same (highlighted yellow) or the third nearest (010) atomic plane (highlighted purple), as depicted in Fig. 1a and b and S5.† Notably, the band structures of all these 20 B/N-codoping configurations are essentially the same, where band-like doping levels are obtained through level crossing between N-pσ (px/pz) and N-pπ (py) doping levels, as shown in Fig. S6.† Such band-like doping levels can be further confirmed by band structure calculations at the HSE06 level38–40 for two arbitrarily selected B/N-codoping configurations (configID-809 and configID-862) among the 20 identified ones, as shown in Fig. 1c and d.
It has also been found that when the B dopant occupies an interstitial site outside the yellow/purple (010) atomic planes found in the 20 identified configurations and the N dopant spatial ordering remains unchanged, which corresponds to another 20 configurations depicted in Fig. S7,† the level crossing between N-pσ and N-pπ doping levels disappears, resulting in multiple localized doping levels, as shown in Fig. S8.† A closer analysis revealed that when the B dopant is located on the yellow/purple (010) atomic planes, a mirror symmetry σ(xz) exists in the 3 × 3 × 1 anatase TiO2 supercell, while it is absent when the B dopant is located elsewhere. Consequently, the symmetries of N-pσ (px/pz) and N-pπ (py) orbitals are different in the first case because of the presence of σ(xz), but are identical in the second case due to the absence of σ(xz). According to group theory,41 when energy levels with different or the same symmetry approach one another, level crossing or level anti-crossing can be expected. This is the reason why level crossing can be observed in the 20 identified configurations (Fig. S6†), while level anti-crossing, giving rise to multiple localized doping levels, has been observed in another 20 configurations (Fig. S8†).
It is important to note that the bandgap narrowing achieved through level crossing between N-pσ/pπ doping levels in the 20 identified B/N-codoping configurations is more significant (∼1.0 eV at the PBE level) than in previously reported N-doped anatase TiO2 (typically less than 0.5 eV at the PBE level18,42). This indicates that the interstitial B dopant can remarkably affect the energies of N doping levels to achieve a more significant bandgap narrowing. To clarify this effect, we took one identified configuration (configID-809) as an example, and compared its band/atomic structures with those of the corresponding N-doping configuration without the interstitial B dopant. As shown in Fig. 2a and d, the interstitial B dopant causes noticeable downshifting/upshifting of two N-pσ doping levels, labeled as N-pσ(1) and N-pσ(2), and the upshifting of N-pσ(2) results in the more significant bandgap narrowing than the case without the interstitial B dopant. By comparing the atomic structures and charge density isosurface plots shown in Fig. 2b and e, the interstitial B dopant substantially shortens the bond length between the Ti atom labeled as Ti3 and its adjacent N dopant, thereby strengthening the corresponding pσ bond. This not only results in the downshifting of N-pσ(1), but also the upshifting of N-pσ(2) as the repulsion between N-pσ(1) and N-pσ(2) is substantially increased.
In order to uncover the reason for the Ti3–N bond shortening, we further analyzed the projected density of states (PDOS). As shown in Fig. 2c and f, the Ti3–N bond shortening can be attributed to the markedly enhanced p–d coupling between N-py (pπ) and Ti3-dyz (t2g) by introducing the interstitial B dopant. This is evident from the more pronounced hybridization peak of these two orbitals and the increased energy of Ti3-dyz in the antibonding region.43 In contrast, the p–d coupling between N-py and Ti-dxy is much weaker as the hybridization peak of the two orbitals is barely changed by introducing the interstitial B dopant, indicating that the extra electron contributed by the interstitial B dopant for compensating the N3−/O2− charge imbalance24,28 preferentially occupies the molecular orbital formed between N-py and Ti-dyz, rather than the one between N-py and Ti-dxy. This selective occupation leads to the shortening of the Ti3–N bond, thereby achieving the significant bandgap narrowing in the identified B/N-codoping configuration.
In order to understand the origin of the much stronger p–d coupling between py and dyz than that between py and dxy, one needs to move from a molecular picture to the crystal. Due to the translational symmetry of a crystal, an atomic orbital, for instance py, at position r should be expressed as a Bloch function ψ = eikRpy, where R denotes a lattice vector and k represents a point in the reciprocal space. In the Bloch function, the k-point determines the phase change of the atomic orbital upon translation between unit cells, and each k-point transforms under a specific point group due to the symmetry of the crystal.41 To be specific, low-symmetry k-points, i.e., those whose point group has fewer symmetry elements, result in the atomic orbital having amplitudes smaller than the maximum value in some unit cells, while high-symmetry k-points ensure that the atomic orbital has the maximum amplitudes in each unit cell.44 Therefore, given a high-symmetry k-point, if two atomic orbitals, for instance py and dyz, transform like the same irreducible representations of its associated point group,45 the high-symmetry k-point generates Bloch function ψ = eikR
(py + dyz), which exhibits strong coupling between the two atomic orbitals within each unit cell.
Based on the above analysis, the much stronger p–d coupling between py and dyz observed in the 20 identified B/N-codoping configurations indicates the existence of high-symmetry k-points that allow a strong p–d coupling in anatase TiO2. In order to validate this speculation, we summarized the point groups of all high-symmetry k-points in the first Brillouin zone of anatase TiO2 in Table S1.† From the table, the point group of the high-symmetry k-point X is C2v, which allows the p–d coupling between py and dyz according to its character table. As illustrated in Fig. 3a and b, under the C2v point group, both py and dyz transform like the same irreducible representation B1, thus allowing strong p–d coupling between these two orbitals.41
In contrast, px/py and dxy transform like different irreducible representations under the point groups associated with all high-symmetry k-points of anatase TiO2. Taking C4v as an example, in this case, (px, py) constitutes a two-dimensional irreducible representation E, while dxy transforms like one-dimensional irreducible representation B2, as illustrated in Fig. 3d and e. As a result, the p–d coupling between py and dxy is much weaker than that between py and dyz, well consistent with the PDOS results shown in Fig. 2c and f. In order to clearly visualize such a difference, we illustrated the charge density isosurface plot of the N doping level above the VBM of TiO2, which is induced by a single N dopant in a 2 × 2 × 1 supercell of anatase TiO2. As shown in Fig. 3c and f, the N doping level mainly originates from the p–d coupling between N-py and Ti-dyz rather than between N-py and Ti-dxy, further confirming the strong coupling between N-py and Ti-dyz.
The existence of strong p–d coupling indicates that the N-py (pπ) are not isolated, since they can strongly couple with O-pπ and Ti-dyz (one of three Ti-t2g orbitals). In this case, a uniform distribution of the N dopants throughout the crystal supercell can ensure a similar N–N distance between different N dopants and prevent them from clustering, thus inducing delocalized molecular orbitals due to the strong coupling between N/O-pπ and Ti-t2g orbitals. As these delocalized molecular orbitals correspond to the band-like doping levels above the VBM of pristine anatase TiO2, the uniform N-doping can be intuitively expected for realizing band-like doping levels in N-doped anatase TiO2. In order to confirm this, we constructed two B/N-codoping configurations with uniform N dopant spatial orderings (Fig. 4a and b) and calculated their band structures at the HSE06 level.38–40 As shown in Fig. 4c and d, band-like doping levels with strong dispersion and apparent overlap with the VBM of pristine anatase TiO2 can be observed for both of the two B/N-codoping configurations, thus achieving a significant bandgap narrowing as much as 0.70 eV.
Based on the above discussions, the existence of strong p–d coupling in anatase TiO2 is essential for achieving band-like doping levels. Due to the strong p–d coupling, the band-like doping levels can be achieved by dopant spatial orderings that are either uniform, or exhibit distinctive periodic features to enable apparent level crossing between N-pσ and N-pπ doping levels. Both of these characteristic dopant spatial orderings match the experimental results, where both the substitutional N and interstitial B contents exhibit very small fluctuations throughout the TiO2 microsphere.25 Notably, as the existence of strong p–d coupling is determined by the crystal symmetry of anatase TiO2, it enables us to effectively break out of the paradigm of correlating the origin of band-like doping levels with the electronic structures of a few specific B/N-codoping configurations. More significantly, it also enables the generalization of this p–d coupling criterion for the prediction of the feasibility of inducing band-like doping levels in other metal oxides. In order to validate this generalization, we further considered the other two polymorphs of TiO2, namely rutile and brookite TiO2.
As for rutile TiO2, the point groups of the high-symmetry k-points in the first Brillouin zone are either D4h or D2h as summarized in Table S1.† Since pπ and t2g orbitals transform like different irreducible representations in both of the point groups, the p–d coupling is much weaker in rutile TiO2 compared to that in anatase TiO2. Aiming to illustrate such a difference between rutile and anatase TiO2, we have calculated the PDOS onto N-pπ and Ti-t2g orbitals for both N-doped anatase and rutile TiO2. As shown in Fig. S9,† the hybridization peak between N-pπ and Ti-t2g orbitals is sharper and has a much higher value in N-doped anatase TiO2 than in N-doped rutile TiO2, demonstrating a much stronger p–d coupling between N-pπ and Ti-t2g orbitals in anatase than in rutile TiO2.
Therefore, it can be predicted that band-like doping levels are hard to be achieved in rutile TiO2. In order to verify this, we also predicted the bandgap values of all possible B/N-codoping configurations in a 2 × 2 × 3 rutile TiO2 supercell by using the established ML method. Specifically, by substituting three lattice O atoms with three N atoms and introducing one interstitial B atom, a total of 2576 symmetrically inequivalent B/N-codoping configurations can be obtained. Among these, 678 representative ones were uniformly sampled for DFT calculations, thus enabling the training of a highly accurate XGBoost model for the prediction of bandgap values, as summarized in Fig. S10.† Based on the ML prediction, we identified the 40 B/N-codoping configurations that exhibit the most significant bandgap narrowing, as shown in Fig. S11.† Besides, we also calculated the band structures of the two B/N-codoping configurations with the most significant bandgap narrowing at the HSE06 level.38–40 As shown in Fig. S12,† the doping levels are largely localized in B/N-codoped rutile TiO2, which is in accordance with the prediction made based on the p–d coupling criterion. Actually, this is in accordance with the previous DFT study on the effect of N doping on rutile TiO2, where N doping typically induces localized doping levels above the VBM of rutile TiO2 according to the PDOS plots.46 Moreover, we also validated the prediction by synthesizing the B/N-codoped rutile TiO2 using an approach similar to that for synthesizing the B/N-codoped anatase TiO2.24,47 As shown in Fig. S13,† only a shoulder-like absorption edge around 450 nm can be observed in the obtained B/N-codoped rutile TiO2. This is typically induced by localized doping levels, since the localized doping levels are discrete within the bandgap, thereby only promoting light absorption at specific wavelengths rather than inducing a band-to-band redshift of the absorption edge. It should also be noted that the obtained B/N-codoped rutile TiO2 has a greenish color rather than red, and a tail-like visible light absorption band can be observed in the UV-visible absorption spectra, which can be ascribed to the existence of a high concentration of oxygen vacancies.48 However, in the N doped rutile TiO2 where substitutional fluorine (F) was adopted as the charge compensator rather than the interstitial B, the obtained N/F-codoped rutile TiO2 has an orange color,49 indicating that the substitutional F can more effectively compensate for the charge imbalance between N3− and O2−, thereby substantially lowering the concentration of oxygen vacancies. Noting that, the N/F-codoping still induces localized doping levels above the VBM of rutile TiO2, which is also in agreement with the prediction made based on the p–d coupling criterion.
For brookite TiO2, its high-symmetry k-point S in the first Brillouin zone allows p–d coupling, as summarized in Table S1.† Thus, it can be predicted that band-like doping levels can be achieved by uniform N doping in brookite TiO2. In order to validate this, by introducing one interstitial B dopant and three substitutional N dopants in a 2 × 2 × 1 brookite TiO2 supercell, we constructed a B/N-codoping configuration with uniform N dopant spatial ordering (Fig. S12c†). As shown in Fig. S12f,† band-like doping levels with strong dispersion and apparent overlap with the VBM of pristine brookite TiO2 can be observed for this B/N-codoping configuration, thus achieving a significant bandgap narrowing as much as 0.63 eV.
Since the pπ orbitals originate from sp2 hybridization of the three-coordinated N/O atoms, and the t2g orbitals arise from the splitting of metal d orbitals by the Oh point group symmetry of the MO6 cluster, rather than being unique to Ti atom. Therefore, the physical picture of the p–d coupling criterion is in principle also applicable for other MO6-cluster-based metal oxide semiconductors. To confirm this, we further studied the effect of N-doping on the electronic structures of four more wide bandgap metal oxides including ScTaO4, WO3, SnO2 and MgTa2O6. By analyzing the point groups as summarized in Table S2,† strong p–d coupling is allowed in ScTaO4 and WO3, while it is absent in both SnO2 and MgTa2O6. Therefore, it can be predicted that band-like doping levels can also be obtained in ScTaO4 and WO3 by uniform N-doping, while this is hard to be fulfilled in SnO2 and MgTa2O6.
In order to validate the prediction for ScTaO4, we constructed a uniform N-doping configuration based on a 2 × 2 × 2 ScTaO4 supercell, and a B/N-codoping configuration by further introducing an interstitial B dopant into the N-doping configuration. The electronic structures of these two doping configurations are shown in Fig. S14.† Clearly, uniform N-doping results in band-like doping levels that merge apparently with the VBM of ScTaO4. Moreover, the charge compensation effect of the interstitial B dopant can further broaden the band-like doping levels to achieve a larger bandgap narrowing of about 1.1 eV. These results well explain the significant overall bandgap narrowing (up to 1.32 eV) achieved experimentally in ScTaO4 by codoping with N and charge compensators.50 As for WO3, we also constructed a uniform N-doping configuration based on a 2 × 2 × 1 WO3 supercell. From the calculated electronic structures shown in Fig. S15,† band-like doping levels above the VBM of WO3 can also be obtained, and this should be the real cause for the smooth shift of the absorption edge towards the visible light region in N-doped WO3 observed in experiment.51
Notably, the demonstrated uniform doping mechanism has significant implications, since it explains, for the first time, why uniform N doping is essential for better performance in metal oxides. This can be highlighted by comparing the first red anatase TiO2 (ref. 24) with the later improved one,25 which features a more uniform N doping instead of gradient N doping. Due to the uniform N doping, not only the band-to-band light absorption but also the photocatalytic oxygen evolution activity was significantly improved. In fact, besides the above metal oxides, uniform N doping is also the key for achieving state-of-the-art photoelectrochemical and photocatalytic performances in other metal oxides, for instance Cs0.68Ti1.83O4,52 where strong p–d coupling exists in both metal oxides according to symmetry analysis, as summarized in Table S3.†
As for the other two metal oxides (SnO2 and MgTa2O6) without strong p–d coupling, the absence of band-like doping levels has been firmly confirmed by considering a sufficient number of N-doping configurations with the most uniform N dopant spatial orderings. In specific, within a 2 × 2 × 3 SnO2 supercell, we first identified the 24 doping configurations with the most uniform distributions of three N dopants, and calculated the corresponding band structures, as shown in Fig. S16.† Clearly, the doping levels are largely localized within the bandgap and exhibit no overlap with the VBM of SnO2. Besides, as for the 24 N-doping configurations with the most uniform distributions of three N dopants within a 2 × 2 × 1 MgTa2O6 supercell, similar localized doping levels can also be observed, as shown in Fig. S17.† These results are in full accordance with the predictions based on the p–d coupling criterion.
Footnote |
† Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d4ta05577b |
This journal is © The Royal Society of Chemistry 2024 |