• Journal of Semiconductors
  • Vol. 40, Issue 9, 092101 (2019)
Xuefen Cai1, Peng Zhang1、2, and Su-Huai Wei1
Author Affiliations
  • 1Beijing Computational Science Research Center, Beijing 100193, China
  • 2College of Physics and Optoelectronic Engineering, Shenzhen University, Shenzhen 518060, China
  • show less
    DOI: 10.1088/1674-4926/40/9/092101 Cite this Article
    Xuefen Cai, Peng Zhang, Su-Huai Wei. Revisit of the band gaps of rutile SnO2 and TiO2: a first-principles study[J]. Journal of Semiconductors, 2019, 40(9): 092101 Copy Citation Text show less

    Abstract

    From the recent experimentally observed conduction band offset and previously reported band gaps, one may deduce that the valence band offset between rutile SnO2 and TiO2 is around 1 eV, with TiO2 having a higher valence band maximum. This implication sharply contradicts the fact that the two compounds have the same rutile structure and the Γ3+ VBM state is mostly an oxygen p state with a small amount of cation d character, thus one would expect that SnO2 and TiO2 should have small valence band offset. If the valence band offset between SnO2 and TiO2 is indeed small, one may question the correctness of the previously reported band gaps of SnO2 and TiO2. In this paper, using first-principles calculations with different levels of computational methods and functionals within the density functional theory, we reinvestigate the long-standing band gap problem for SnO2. Our analysis suggests that the fundamental band gap of SnO2 should be similar to that of TiO2, i.e., around 3.0 eV. This value is significantly smaller than the previously reported value of about 3.6 eV, which can be attributed as the optical band gap of this material. Similar to what has been found in In2O3, the discrepancy between the fundamental and optical gaps of SnO2 can be ascribed to the inversion symmetry of its crystal structure and the resultant dipole-forbidden transitions between its band edges. Our results are consistent with most of the optical and electrical measurements of the band gaps and band offset between SnO2 and TiO2, thus provide new understanding of the band structure and optical properties of SnO2. Experimental tests of our predictions are called for.

    1. Introduction

    Post-transition-metal oxides, such as In2O3 and SnO2, have been extensively used as transparent conducting oxides (TCOs), a unique class of materials that combine the high electrical conductivity and optical transmission in the visible or near-infrared spectral range, which are needed in various optoelectronic devices, such as solar cells, flat-panel displays, light emitting diodes (LEDs) and transparent transistors[1-7]. Obviously, one of the most critical properties of these materials is related to their band gaps that essentially determine their transparency. However, the determination of band gaps of these TCOs has not been a trivial work. For instance, early experimental measurements on In2O3 revealed a strong optical absorption at 3.75 eV and a much weaker absorption at 2.62 eV[8]. Initially, the former absorption was attributed as the direct band gap of In2O3, while the latter was attributed to the indirect transitions. Although this interpretation seems able to explain some of the experiments, the indirect transitions in In2O3 has never been confirmed by first-principles calculations[9, 10], hindering further understanding of the band structure and optical properties of this material. More recently, by combining the experimental measurements and theoretical calculations, Walsh et al.[11] identified that In2O3 actually possesses a direct fundamental gap of 2.9 eV and a much larger optical gap of 3.7 eV. The large discrepancy between the fundamental and optical gaps was attributed to the inversion symmetry of its bixbyite structure and the resultant dipole-forbidden transitions between the valence band maximum (VBM) and conduction band minimum (CBM) states. This argument, capable of explaining various theoretical and experimental results, has now been commonly accepted.

    In contrast to the case of In2O3, the band gap problem of SnO2 is still not fully understood, largely because of the lack of detailed experimental measurements. Analogous to bixbyite In2O3, rutile SnO2 also has the inversion symmetry and the dipole-forbidden transitions between the band edge states. As a result, one may expect different fundamental and optical gaps of SnO2 as well, which, however, has never been reported in experiments. The single-photon and two-photon absorption experiments showed an optical gap of SnO2 at ~3.60 eV, which has been widely taken as its “band gap”[12-14]. Moreover, the recent theoretical calculations, based on the density functional theory (DFT) and GW approximation, identified a fundamental gap of 3.65 eV for SnO2[15, 16], which is nearly equal to its optical gap. Sabino et al.[17] attributed the coincidence between the measured fundamental and optical gaps to the intense illumination used in experiments, so that the very weak absorption in the vicinity of Γ point at the band edges was detected. Based on this argument, they also predicted that a much larger optical band gap of 4.34 eV should be detected, if the low illumination condition is satisfied. However, up to now, no experimental demonstration of this difference between fundamental and optical gaps has been reported.

    Despite the absence of direct measurements on its fundamental gap, more insight can be obtained from the recent experiments where SnO2 was employed as the electron transport layer (ETL) in both dye-sensitized and perovskite solar cells to replace the commonly used TiO2[18, 19]. The band alignment measurements[19] showed that the CBM of SnO2 is about 0.4 eV lower than that of TiO2, which facilitates the electron transport. Assuming the fundamental gaps of SnO2 and TiO2 as 3.65 and 3.03 eV[20], respectively, we can easily deduce that the VBM of SnO2 is ~1.02 eV lower than that of TiO2, as illustrated in Fig. 1(a). This result, however, is quite puzzling and contradicts the small VBM offset one would expect in this system[21-23], because the two compounds are isovalent semiconductors with the same rutile structure and similar lattice constants, and moreover, the Γ3+ VBM state is almost a pure oxygen p state, which only couples weakly with the cation d state but does not couple to the cation p state. Therefore, this VB offset anomaly is difficult to understand if the fundamental gap of SnO2 is taken as ~3.65 eV, as previously reported.

    (Color online) Band gaps and band alignments between rutile SnO2 and TiO2. In (a) the band gaps of SnO2 is taken from Ref. [15] and the CB band alignment is taken from Ref. [19], while in (b) both of them are calculated in this work. The band gap of TiO2 is always set to its experimental value (taken from Ref. [20]).

    Figure 1.(Color online) Band gaps and band alignments between rutile SnO2 and TiO2. In (a) the band gaps of SnO2 is taken from Ref. [15] and the CB band alignment is taken from Ref. [19], while in (b) both of them are calculated in this work. The band gap of TiO2 is always set to its experimental value (taken from Ref. [20]).

    To address this issue, in this work, we revisit the band gap problem for the rutile SnO2 by using first-principles calculations. Different-level computational methods and functionals, such as PBE, PBE + G0W0, HSE06 and HSE06 + G0W0 are employed and the results are carefully compared. We find that the HSE06 calculations yield a VB offset of 0.38 eV between SnO2 and TiO2, generally obeying the common-anion rule. The fundamental gap is calculated to be 2.96 eV, which is much smaller than the previously reported value. The conduction band (CB) offset is then found to be 0.45 eV, in good agreement with experiments[18, 19]. Moreover, similar to In2O3, our calculations reveal that the direct optical transition between the band edges of SnO2 is parity forbidden and the optical gap between the lower-lying VB states and CBM is 3.69 eV, which can well explain its optical absorptions[12-14]. On the other hand, by employing the HSE06 + G0W0 method, the fundamental gap of SnO2 is found to be 3.76 eV that is close to the value reported previously[16, 17, 24]. However, with the same method, the fundamental gap of TiO2 is calculated to be 3.68 eV, significantly larger than the experimental result of 3.03 eV[20]. In fact, we find that the fundamental gap of SnO2 is always similar to that of TiO2, independent of the computational methods and functionals employed, which implies that the HSE06 + G0W0 calculations may overestimate the fundamental gap of SnO2, which has been the case also found in other sp semiconductors[25]. The estimated error of HSE06 + G0W0 calculation is about 0.8 eV, which in turn results in a fundamental gap of SnO2 as ~3.0 eV, in good agreement with the HSE06 calculations.

    2. Theoretical calculation

    Our calculations are carried out using the projector augmented wave (PAW) method[26] and the Heyd–Scuseria–Ernzerhof (HSE06) hybrid functional[27] within the density functional theory (DFT) as implemented in VASP[28]. For both SnO2 and TiO2, a 520 eV cutoff energy is used for the plane wave basis set and a Γ-centered 4 × 4 × 6 k-point sampling of the Brillouin zone is employed. Since the band gaps of these compounds are sensitive to their lattice constants, experimental lattice constants (a = 4.737 Å and c = 3.186 Å for SnO2[29], a = 4.594 Å and c = 2.958 Å for TiO2[30]) are adopted for the electronic structure calculations. During the structural relaxation, the atomic coordination is fully relaxed until the Hellman–Feynman force on each atom is less than 0.01 eV/Å. To compare the influence of different computational methods and functionals on the band gaps, the Perdew–Burke–Ernzerhof (PBE)[31], GW[32] with the PBE (PBE + G0W0), and HSE06 + G0W0 calculations are also performed.

    3. Results and discussion

    Fig. 2(a) illustrates the tetragonal rutile structure for both SnO2 and TiO2 with the space group . As shown in Fig. 2(a), each cation atom (Sn or Ti) binds with six O atoms, forming a distorted octahedron, while each O atom is trigonally coordinated to three cation atoms. The calculated projected density of states (PDOS) are given in Fig. 2(b). We find that the Γ1+ CBM state of SnO2 is mainly composed of Sn 5s and O 2s states, while the Γ3+ VBM state is mainly composed of O 2p states. For TiO2, on the other hand, the CBM has mostly Ti 3d character, while the VBM is still the Γ3+ state dominated by O 2p states with minor Ti 3d character. The above analysis indicates that the well-known common-anion rule[21, 22] should apply to this system, i.e., the VBM offset between these two compounds should be relatively small. To verify this expectation, we construct a SnO2/TiO2 heterojunction consisting of six SnO2(001) and six TiO2(001) layers, as shown in Fig. 2(c), to calculate the VB offset between SnO2 and TiO2 using the method described in Ref. [33], that is

    (Color online) (a) Primitive unit cells and (b) calculated total and partial density of states for rutile SnO2 and TiO2. The green, blue and purple balls represent Sn, Ti and O respectively. (c) Side view of the atomic configuration for SnO2/TiO2 heterojunction (top panel).

    Figure 2.(Color online) (a) Primitive unit cells and (b) calculated total and partial density of states for rutile SnO2 and TiO2. The green, blue and purple balls represent Sn, Ti and O respectively. (c) Side view of the atomic configuration for SnO2/TiO2 heterojunction (top panel).

    $ \Delta {E_\rm{V}}(\rm{SnO}_\rm{2}/\rm{TiO}_\rm{2}) = {({E_\rm{V}} - {E_\rm{CL}})_{\rm{TiO}_\rm{2}}}} - {({E_\rm{V}} - {E_\rm{CL}})_{\rm{SnO}_\rm{2}}} + \Delta {E_\rm{CL}}. $  (1)

    The first two terms on the right-hand side are the core-level to VBM energy separations for pure bulk TiO2 and SnO2, respectively, and the third term is the difference of core-level binding energy between SnO2 and TiO2 in the SnO2/TiO2 superlattice. Since the lattice mismatch between SnO2(001) and TiO2(001) layers is very small (< 2%), here we set the lattice constants of the superlattice as the average of those of SnO2 and TiO2 and fully relax the atomic coordination in the superlattice with a Γ-centered 4 × 4 × 1 k-mesh until the Hellman–Feynman force on each atom is less than 0.01 eV/Å.

    The calculated band alignment for SnO2 and TiO2 is shown in Fig. 1(b). We find that the VBM of SnO2 is 0.38 eV lower in energy than that of TiO2, which generally follows the common-anion rule. The slightly lower VBM of SnO2 can be attributed to the more polar Sn4+ ions, compared to Ti4+ ions. The calculated fundamental gaps of SnO2 and TiO2 are 2.96 and 3.15 eV, respectively, with the latter slightly larger than the experimental value of 3.03 eV[20]. Using the experimental band gap of TiO2, one can deduce that the CBM of SnO2 lies ~0.45 eV below that of TiO2, which falls in line with the experimental measurements[18, 19]. It is interesting to see that the calculated fundamental gap of SnO2 is slightly smaller than that of TiO2, which is quite different from the experimentally measured band (optical) gap of ~3.60 eV[12-14]. To understand this discrepancy, we carefully investigate the band structure of SnO2, as shown in Fig. 3(a). Apparently, SnO2 has a direct fundamental gap located at the Γ point. Inspection of the wavefunction at the band edges indicates that the VBM and CBM states both have the even parity with Γ3+ and Γ1+ symmetries, respectively. Therefore, the electric-dipole transition between these two states is forbidden. Table 1 gives the corresponding transition matrix elements. We can see that the transition probability between the VBM and CBM at the Γ point is exactly zero. As the k-point is slightly off the Γ-point, the transition matrix elements between the VB and CB states are still negligibly small, indicating very weak optical transitions in this region. In contrast, we find that the pronounced optical transitions can occur from the low-lying VB states to the CBM at the Γ point, leading to an optical gap 0.74 eV larger than the fundamental gap. As shown in Fig. 3(a), the low-lying VB states have the Γ5 symmetry with oxygen and cation p characters, so the optical transitions from these states to the CBM state are allowed. As shown in Fig. 3(b), the calculated absorption coefficient indicates a very slow increase above 2.96 eV and a sharp increase above 3.70 eV, in good agreement with the band structure analysis. From these calculations, we can conclude that SnO2 should have a relatively small fundamental gap of ~3.0 eV and a much larger optical gap of ~3.7 eV, with the latter accounting for the “band gap” observed in optical absorption experiments[12-14].

    (Color online) (a) Band structure of SnO2 along two high symmetry lines M–Γ–Z. (b) Optical absorption coefficient of SnO2. The valence band maximum is set at zero. In (a), the red and green arrows represent the possible optical transitions at Γ-point and away from Γ-point, respectively.

    Figure 3.(Color online) (a) Band structure of SnO2 along two high symmetry lines M–Γ–Z. (b) Optical absorption coefficient of SnO2. The valence band maximum is set at zero. In (a), the red and green arrows represent the possible optical transitions at Γ-point and away from Γ-point, respectively.

    Table Infomation Is Not Enable

    It should be noted that the fundamental gap of SnO2 obtained here is quite different from the values reported by the previous calculations[15-17]. For instance, based on the HSE03 + G0W0 calculations, Schleife et al.[15] reported a fundamental gap of SnO2 as 3.65 eV that is similar to observed optical gap. They attributed the near coincidence between the fundamental and optical gaps to the dipole-allowed transitions in the vicinity of the Γ point, as depicted by the green arrow in Fig. 3. Sabino et al.[17] further explained that the detection of these weak transitions can be attributed to the intense illumination used in experiments. To compare with these previous calculations, we also employ the GW formalism combined with the hybrid functional calculations (HSE06 + G0W0), together with some other methods, such as PBE and PBE + G0W0, with the calculated fundamental gaps of SnO2 and TiO2 given in Table 2. For the HSE06 + G0W0 calculations, the fundamental gap of SnO2 is found to be 3.76 eV that is consistent with the previous calculations. However, with the same method, the fundamental gap of TiO2 is calculated to be 3.68 eV, apparently larger than the experimental results[20], which indicates that the HSE06 + G0W0 calculations cannot describe the band structure well, at least for TiO2. More interestingly, we can see from Table 2 that the calculated fundamental band gap of SnO2 is always similar to that of TiO2, although different computational methods and/or functionals result in different band gap values. This trend suggests that the HSE06 + G0W0 calculations, as employed in the previous work, may also overestimate the fundamental gap of SnO2, which has also been the case found in other sp semiconductors[25].

    Table Infomation Is Not Enable

    One may argue that the hybrid functional and/or GW approximation may have different impacts on the fundamental gaps of SnO2 and TiO2, because of the different component of their CBM states, as shown in Fig. 1(b). To clarify this point, we consider a series of metal oxides, including ZnO, CdO, Ga2O3, In2O3 and GeO2, for which the CBM and VBM components are all similar to that of SnO2. Table 3 gives the calculated fundamental gaps by the HSE06 + G0W0 method, together with the available experimental data. We find that although the HSE06 + G0W0 calculations can result in reasonable fundamental gaps for IIB oxides ZnO and CdO, which have large d orbital component at the VBM, it significantly overestimates the band gaps of IIIA oxides Ga2O3, In2O3 and IVA oxides GeO2, which have similar band edge wavefunction characters as SnO2. We find the differences between the calculated and experimentally measured fundamental gaps increase monotonically as the occupied d orbital binding energy increases, e.g., from ZnO to Ga2O3 and to GeO2 or from CdO to In2O3, as summarized in Fig. 4. From this trend, it is reasonable to expect that the HSE06 + G0W0 calculations would also overestimate the fundamental gap of SnO2. By a simple linear extrapolation, we can roughly estimate the overestimation is ~0.8 eV, which gives an error-corrected fundamental gap of SnO2 as ~3.0 eV, in accordance with the HSE06 calculations.

    Table Infomation Is Not Enable

    (Color online) Differences between the HSE06 + G0W0 calculated and experimentally measured band gaps for a series of oxides.

    Figure 4.(Color online) Differences between the HSE06 + G0W0 calculated and experimentally measured band gaps for a series of oxides.

    Recently, a hard X-ray photoelectron spectroscopy (HAXPES) measurement showed that the sharp increase of the VBM occurs at ~3.6 eV below the Fermi energy for undoped SnO2[39]. To confirm that this is consistent with our prediction that the fundamental band is only 2.9 eV, we have plotted the total DOS of SnO2 with high resolution, as shown in Fig. 5. We see that the sharp increase of the band edge DOS occurs at 0.7 eV below the VBM, corresponding to the A points in the band structure (see Fig. 3), that is, it gives also the optical band gap, not the fundamental band gap. This is consistent with the fact that due to the specific band structure of SnO2 (Fig. 3), the HAXPES signal near the VBM is low, so can be easily missed by the HAXPES measurement by weak signal.

    (Color online) Calculated DOS of SnO2 for the valence bands below VBM level. The right-hand side shows the DOS with enlarged energy scale below the VBM at E = 0 eV.

    Figure 5.(Color online) Calculated DOS of SnO2 for the valence bands below VBM level. The right-hand side shows the DOS with enlarged energy scale below the VBM at E = 0 eV.

    4. Conclusion

    In conclusion, we have performed a systematic first-principles calculations and symmetry analysis to revisit the long-standing band gap problem for SnO2. Different-level computational methods and functionals, such as PBE, PBE + G0W0, HSE06 and HSE06 + G0W0 are employed. We find that in all these calculations the band gap of SnO2 is similar or slightly smaller than that of TiO2. The HSE06 calculations yield a fundamental gap of 3.0 eV and a much larger optical gap of 3.6 eV for SnO2, which is consistent with recent experimental measurement of the band alignment between SnO2 and TiO2, but smaller than the previously reported fundamental band gap of 3.65 eV for SnO2. The discrepancy between the fundamental and optical gaps is attributed to the inversion symmetry of rutile SnO2 and the resultant dipole-forbidden transitions between the VBM and CBM, similar to what has been found in In2O3. On the other hand, we find that the HSE06 + G0W0 method, as employed in previous calculations, would overestimate the fundamental gaps of both TiO2 and SnO2 as well as a series of other oxides, including Ga2O3, In2O3 and GeO2. More experimental tests of our predictions are called for. Because SnO2 is widely used in optoelectronic devices, our new understanding of the band structure and optical properties of SnO2 should have large impact on the future design of optoelectronic materials.

    Acknowledgment

    We acknowledge the computational support from the Beijing Computational Science Research Center (CSRC). This work was supported by the Science Challenge Project (No. TZ2016003), the National Key Research and Development Program of China (No. 2016YFB0700700), and the Nature Science Foundation of China (No. 11634003, 51672023, U1930402 ).

    References

    [1] K Chopra, S Major, D Pandya. Transparent conductors-a status review. Thin Solid Films, 102, 1(1983).

    [2] K Nomura, H Ohta, K Ueda et al. Thin-film transistor fabricated in single-crystalline transparent oxide semiconductor. Science, 300, 1269(2003).

    [3] J F Wager. Transparent electronics. Science, 300, 1245(2003).

    [4] K Nomura, H Ohta, A Takagi et al. Room-temperature fabrication of transparent flexible thin-film transistors using amorphous oxide semiconductors. Nature, 432, 488(2004).

    [5] T Minami. Transparent conducting oxide semiconductors for transparent electrodes. Semicond Sci Technol, 20, S35(2005).

    [6] C G Granqvist. Transparent conductors as solar energy materials: a panoramic review. Sol Energy Mater Sol Cells, 91, 1529(2007).

    [7] K H Zhang, K Xi, M G Blamire et al. P-type transparent conducting oxides. J Phys Condens Matter, 28, 383002(2016).

    [8] R Weiher, R Ley. Optical properties of indium oxide. J Appl Phys, 37, 299(1966).

    [9] P Erhart, A Klein, R G Egdell et al. Band structure of indium oxide: Indirect versus direct band gap. Phys Rev B, 75, 153205(2007).

    [10] S Z Karazhanov, P Ravindran, P Vajeeston et al. Phase stability, electronic structure, and optical properties of indium oxide polytypes. Phys Rev B, 76, 075129(2007).

    [11] A Walsh, J L Da Silva, S H Wei et al. Nature of the band gap of In2O3 revealed by first-principles calculations and X-ray spectroscopy. Phys Rev Lett, 100, 167402(2008).

    [12] K Reimann, M Steube. Experimental determination of the electronic band structure of SnO2. Solid State Commun, 105, 649(1998).

    [13] D Fröhlich, R Kenklies, R Helbig. Band-gap assignment in SnO2 by two-photon spectroscopy. Phys Rev Lett, 41, 1750(1978).

    [14] M Nagasawa, S Shionoya. Temperature dependence of the fundamental optical absorption edge in stannic oxide. J Phys Soc Jpn, 30, 1118(1971).

    [15] A Schleife, J Varley, F Fuchs et al. Tin dioxide from first principles: Quasiparticle electronic states and optical properties. Phys Rev B, 83, 035116(2011).

    [16] J Berger, L Reining, F Sottile. Efficient GW calculations for SnO2, ZnO, and rubrene: The effective-energy technique. Phys Rev B, 85, 085126(2012).

    [17] F P Sabino, L N Oliveira, i S H Wei et al. Optical and fundamental band gaps disparity in transparent conducting oxides: new findings for the In2O3 and SnO2 systems. J Phys: Condens Matter, 29, 085501(2017).

    [18] H J Snaith, C Ducati. SnO2-based dye-sensitized hybrid solar cells exhibiting near unity absorbed photon-to-electron conversion efficiency. Nano Lett, 10, 1259(2010).

    [19] Q Jiang, L Zhang, H Wang et al. Enhanced electron extraction using SnO2 for high-efficiency planar-structure HC(NH2)2PbI3-based perovskite solar cells. Nat Energy, 2, 16177(2017).

    [20] D O Scanlon, C W Dunnill, J Buckeridge et al. Band alignment of rutile and anatase TiO2. Nat Mater, 12, 798(2013).

    [21] W A Harrison. Elementary theory of heterojunctions. J Vac Sci Technol, 14, 1016(1977).

    [22] S H Wei, A Zunger. Role ofdorbitals in valence-band offsets of common-anion semiconductors. Phys Rev Lett, 59, 144(1987).

    [23] S H Wei, A Zunger. Calculated natural band offsets of all II–VI and III–V semiconductors: Chemical trends and the role of cation d orbitals. Appl Phys Lett, 72, 2011(1998).

    [24] A Walsh, J L F Da Silva, S H Wei. Multi-component transparent conducting oxides: Progress in materials modelling. J Phys: Condens Matter, 23, 334210(2011).

    [25] W Chen, A Pasquarello. Band-edge positions in GW: effects of starting point and self-consistency. Phys Rev B, 90, 165133(2014).

    [26] G Kresse, D Joubert. From ultrasoft pseudopotentials to the projector augmented-wave method. Phys Rev B, 59, 1758(1999).

    [27] J Heyd, G E Scuseria, M Ernzerhof. Hybrid functionals based on a screened Coulomb potential. J Chem Phys, 118, 8207(2003).

    [28] G Kresse, J Furthmüller. Efficiency of ab-initio total energy calculations for metals and semiconductors using a plane-wave basis set. Comput Mater Sci, 6, 15(1996).

    [29] J Haines, J M Léger. X-ray diffraction study of the phase transitions and structural evolution of tin dioxide at high pressure: Relationships between structure types and implications for other rutile-type dioxides. Phys Rev B, 55, 11144(1997).

    [30] M Landmann, E Rauls, W G Schmidt. The electronic structure and optical response of rutile, anatase and brookite TiO2. J Phys: Condens Matter, 24, 195503(2012).

    [31] J P Perdew, K Burke, M Ernzerhof. Generalized gradient approximation made simple. Phys Rev Lett, 77, 3865(1996).

    [32] F Aryasetiawan, O Gunnarsson. The GW method. Rep Prog Phys, 61, 237(1998).

    [33] S H Wei, A Zunger. Valence band splittings and band offsets of AlN, GaN, and InN. Appl Phys Lett, 69, 2719(1996).

    [34] Y Kang, G Kang, H H Nahm et al. GW calculations on post-transition-metal oxides. Phys Rev B, 89(2014).

    [35] D C Reynolds, D C Look, B Jogai et al. Valence-band ordering in ZnO. Phys Rev B, 60, 2340(1999).

    [36] F P Koffyberg. Thermoreflectance spectra of CdO: Band gaps and band-population effects. Phys Rev B, 13, 4470(1976).

    [37] M Rebien, W Henrion, M Hong et al. Optical properties of gallium oxide thin films. Appl Phys Lett, 81, 250(2002).

    [38] A Trukhin, M Kink, Y Maksimov et al. Luminescence of GeO2 glass, rutile-like and α-quartz-like crystals. J Non-Cryst Solids, 352, 160(2006).

    [39] F Borgatti, J A Berger, D Céolin et al. Revisiting the origin of satellites in core-level photoemission of transparent conducting oxides: The case of n-doped SnO2. Phys Rev B, 97, 155102(2018).

    Xuefen Cai, Peng Zhang, Su-Huai Wei. Revisit of the band gaps of rutile SnO2 and TiO2: a first-principles study[J]. Journal of Semiconductors, 2019, 40(9): 092101
    Download Citation