Nature of the Band Gap of In2O3 Revealed by First-Principles Calculations and X-Ray Spectroscopy Aron Walsh,1,* Juarez L. F. Da Silva,1 Su-Huai Wei,1 C. Körber,2 A. Klein,2 L. F. J. Piper,3 Alex DeMasi,3 Kevin E. Smith,3 G. Panaccione,4 P. Torelli,5 D. J. Payne,6 A. Bourlange,6 and R. G. Egdell6 1National Renewable Energy Laboratory, Golden, Colorado 80401, USA 2Darmstadt University of Technology, 64287 Darmstadt, Germany 3Department of Physics, Boston University, 590 Commonwealth Avenue, Boston, Massachusetts 02215, USA 4Laboratorio TASC, INFM-CNR, Area Science Park, S.S. 14, Km 163.5, 34012 Trieste, Italy 5CNR-INFM-S3, Via Campi 213/A, I-41100 Modena, Italy 6Chemistry Research Laboratory, Mansfield Road, Oxford OX1 3TA, United Kingdom (Received 5 November 2007; published 25 April 2008) Bulk and surface sensitive x-ray spectroscopic techniques are applied in tandem to show that the valence band edge for In2O3 is found significantly closer to the bottom of the conduction band than expected on the basis of the widely quoted bulk band gap of 3.75 eV. First-principles theory shows that the upper valence bands of In2O3 exhibit a small dispersion and the conduction band minimum is positioned at �. However, direct optical transitions give a minimal dipole intensity until 0.8 eV below the valence band maximum. The results set an upper limit on the fundamental band gap of 2.9 eV. DOI: 10.1103/PhysRevLett.100.167402 PACS numbers: 78.70.En, 73.20.At, 78.20.Bh Despite the widespread use of In2O3 as a transparent contact in photovoltaic devices, liquid crystal displays, and light emitting diodes [1,2], the nature of the band gap in this material remains contentious [3–6]. Early measure- ments on In2O3 single crystals showed that the onset of strong optical absorption is found to be 3.75 eV, but with indications of a much weaker absorption onset at 2.62 eV attributed to indirect electronic transitions [7]. Nonetheless the ‘‘band gap’’ of In2O3 is widely quoted as 3.75 eV [8] or thereabouts [9]. However, the top of the valence band in x- ray photoemission spectra of lightly doped n-type In2O3 is found to be less than 3 eV below the Fermi level (which is situated just above the bottom of the conduction band), adding further weight to the hypothesis that In2O3 may have an indirect band gap [10]. It has been argued [5] that an indirect gap could arise from mixing of shallow core In 4d states with O 2p states away from the � point within the centrosymmetric crystal structure of In2O3: a similar situ- ation pertains in rocksalt CdO where the lowest energy gap is undoubtedly indirect [5,11]. However, band structure calculations [3] on In2O3 have consistently failed to find upward dispersion of the topmost valence band by the 1 eV required by the indirect gap hypothesis. It has therefore been argued that the onset of valence band photoemission intensity is reduced from the value expected on the basis of the bulk band gap by upward band bending at the surface and that the weak optical absorption around 2.6 eV is associated with defects [4,6]. In this Letter we compare the positions of valence band edges in conventional x-ray photoemission spectra with those measured with the very much less surface sensitive techniques of hard x-ray photoemission and x-ray emission spectroscopies (XPS and XES): while the valence electron inelastic mean free path length in Al K� XPS is probably of the order of 25 Å, under 6000 eV excitation the path length increases to values around 60 Å [12]. XES is a photon-in photon-out technique with an effective sampling depth of order 1000 Å. We find that there is no significant shift in spectral features between the different experimen- tal techniques, which argues strongly against the band bending model. Moreover we further demonstrate, through symmetry analysis of the band structure, that direct optical transitions at � from the valence band maximum (VBM) to the conduction band minimum (CBM) are parity forbidden and that the first strong transitions occur from valence bands 0.81 eV below the VBM. This new insight solves the long-standing ‘‘band gap discrepancy’’ and calls for a reinterpretation of all experimental and theoretical studies of In2O3 that require knowledge of the fundamental gap. In2O3 and Sn-doped In2O3 (2% and 10 weight % Sn) were deposited by radio frequency magnetron sputtering onto doped Si substrates to a thickness of 400–500 nm using pure Ar as the sputter gas and a 400 �C substrate temperature. Carrier concentrations were estimated from conductivity values using a mobility of 30 cm2 V�1 s�1 and from measured energies of plasmon satellites on In 3d core lines [10]. For the most highly doped sample the carrier concentration reaches a value of 1:2� 1021 cm�3. Conventional x-ray photoemission spectra were measured in situ in the deposition system with a Physical Electronics Phi 5700 spectrometer and ex situ with monochromatic Al K� radiation (h� � 1486:6 eV) in both a Scienta XPS system (incorporating a rotating anode Al K� x-ray source and a 300 mm mean radius electron energy analyzer) with an overall energy resolution of 0.35 eV and a VG ESCALAB 250. Energies were referenced relative to Fermi level onsets of Ag samples used to calibrate the spectrometers. The silver Fermi level was found to coin- cide with weak but well-defined Fermi edges found for the doped In2O3 samples. Further Al K� measurements were PRL 100, 167402 (2008) P H Y S I C A L R E V I E W L E T T E R S week ending 25 APRIL 2008 0031-9007=08=100(16)=167402(4) 167402-1 © 2008 The American Physical Society Urheberrechtlich geschützt / In Copyright http://dx.doi.org/10.1103/PhysRevLett.100.167402 https://rightsstatements.org/page/InC/1.0/ performed on (001) oriented single crystal thin film samples grown by O-plasma assisted molecular beam epi- taxy on Y-doped ZrO2 substrates. The positions of the valence band onsets were consistent between all the differ- ent data sets. Hard x-ray photoemission spectra (HXPS) at h� � 6000 eV were measured on beam line 16 at the ESRF [12] again with an overall energy resolution of 0.35 eV. Here spectra were referenced to the Fermi ener- gies of the Sn-doped samples. Finally x-ray emission spec- tra at the O K edge were measured at the soft x-ray undulator beam line X1B at the National Synchrotron Light Source (NSLS), Brookhaven National Laboratory. The emission spectra were obtained using a Nordgren-type grazing incidence spherical grating spectrometer with an energy resolution set to 0.35 eV at the O K edge [13]. The incident photon energy was set far above the absorption threshold at 567 eV. The emission energy axes were cali- brated to the second order L edge of a zinc metal, for the same settings. The experimental data are shown in Fig. 1. In both Al K�XPS and HXPS well-defined structure is found close to the Fermi energy which grows in intensity with increasing doping and extends down to about 1.5 eV binding energy in the most highly doped sample. This is associated with occupied conduction band states. The intensity of the conduction band feature increases dramatically relative to the intensity of the valence band edge in switching from excitation at h� � 1486:6 eV to h� � 6000 eV due to the pronounced In 5s character of the conduction band states and the increasing value of the cross section for In 5s states relative to that for O 2p states with increasing photon energy [14]. The valence band edge is seen to lie 2.9 eV below the Fermi level in both Al K� XPS and HXPS for the nominally undoped sample and moves to higher bind- ing energies of 3.2 and 3.5 eV, respectively, for 2% and 10% doped samples. All binding energies are referenced to the Fermi level and the shift arises from the upward move- ment of the Fermi level within the conduction band with increasing doping partially offset by band gap shrinkage with doping. Conduction band occupancy produces the well known Burstein-Moss shift in optical absorption [8]. The magnitudes of the observed band edge shifts are con- sistent with the known free carrier concentrations [8]. The valence band onset energy of 2.9 eV for the nominally undoped sample sets an upper limit for the fundamental band gap of In2O3 assuming little or no surface band bending: this value is further reduced by the occupation of conduction band states arising from adventitious donor defects. A distinct Fermi edge onset was also observed in the x- ray emission spectrum (which is determined by the bulk occupied O 2p like partial density of states) of the 10% Sn- doped sample at h� � 530:06 eV with a sigmoidal width matching the spectrometer resolution of 0.35 eV. This feature was absent in the spectrum of the nominally un- doped sample and allows the x-ray emission data to be transposed onto the same binding energy scale as the photoemission data. Although strong band tailing prevents accurate determination of the valence band onset in this spectrum, the valence band peak maximum relative to the Fermi energy is seen to coincide with that in Al K� XPS. Thus the binding energy scales for all three techniques coincide, a fact which argues strongly against the band bending model. Similar to Tl2O3 [15], In2O3 adopts the body centered cubic bixbyite (FeMnO3) structure (space group Ia�3, T7 h symmetry) with 8 f.u. per primitive cell [16]. Each In atom is coordinated by six oxygen atoms in a distorted octahe- dron, with In-O interatomic distances ranging from 2.13– 2.23 Å and a lattice constant of 10.12 Å [16]. The equilib- rium geometric and electronic structure of In2O3 were calculated using density functional theory [17] within the generalized-gradient approximation [18] (GGA-PBE) and the all electron projector-augmented-wave method [19] as implemented in VASP [20,21]. A plane wave cutoff of FIG. 1 (color online). (a) Al K� XPS (h� � 1486:6 eV) of nominally undoped and 2% and 10% Sn-doped In2O3 thin films. (b) HXPS (h� � 6000 eV) of the same films. (c) Al K� XPS and O K shell XES of 10% Sn-doped In2O3 aligned using the Fermi edge (0 eV) visible in both spectra. PRL 100, 167402 (2008) P H Y S I C A L R E V I E W L E T T E R S week ending 25 APRIL 2008 167402-2 400 eV and a 4� 4� 4 k-point mesh were found to give sufficient convergence. The calculated band structure along two high symmetry linesH�1=2;�1=2; 1=2�-��0; 0; 0�-N�0; 0; 1=2� is shown in Fig. 2. The highest occupied band exhibits very little dispersion, e.g., 19 meV along the �-N line. This is clearly inconsistent with the existence of a strongly indirect fun- damental gap in the order of 1 eV. The effect of electron correlation on the In 4d states was investigated by using GGA�U (U� J � 5 eV). The effect is small, resulting in a change of �-N band width of less than 50 meV. The VBM state at � is threefold degenerate and is derived from O 2p and In 4d character (�4, Tg symmetry), while the CBM state is a mixture of In 5s and O 2s orbitals (�1, Ag symmetry). As bixbyite contains an inversion center and the electric-dipole operator is of odd parity, strong optical transitions are only permitted between two states of opposing parity. These symmetry require- ments result in a zero optical transition matrix element for direct VBM to CBM absorption at �, confirming that this �4-�1 transition is formally forbidden and can only make a very weak contribution to photon absorption under the influence of lattice vibrations. It is only from 0.81 eV below the VBM that strong transitions are observed. Here, the wave function character at � becomes sufficiently p like (�8, Tu symmetry) and strong optical transitions can occur, with a transition matrix element of 0.74 a.u. for �8-�1 absorption. Five sets of bands lie in the range between 0–0.81 eV, of which three are of even parity, while two are of odd parity; however, while transitions from the odd parity bands at 0.25 and 0.50 eV are nonzero, they result in very small matrix elements on the order of 100 times weaker than the strong �8-�1 absorption at 0.81 eV. This is because these states are derived from the ‘‘fold-in’’ bands of the ideal In2O3 primitive cell, so they are only pseudodirect. Therefore, both the computed opti- cal matrix elements and the resulting absorption coefficient (Fig. 3) are consistent with a fundamental gap 0.81 eV lower in energy than the optical band gap. While the absorption spectrum, summed over all possible direct tran- sitions from a dense sampling of the Brillouin zone, is calculated with a VBM-CBM separation of 2.89 eV [22], the onset of optical absorption is only observed at 3.70 eV, after which the intensity rises with increasing photon en- ergy. This demonstrates that the valence band states within 0.81 eV of the VBM make no significant contribution to low energy photon absorption in the frozen bulk crystal; weak transitions can result experimentally through local symmetry breaking. In2O3 is the prototypical n-type transparent conducting oxide [1]. Occupation of the conduction band through n-type doping, as observed from the increased Fermi en- ergy–VBM separation in the experimental spectra on in- troduction of Sn (Fig. 1), can induce transitions away from the � point due to a Burstein-Moss shift. To take this effect into account theoretically, we have examined the magni- tude of the direct optical transitions along the H-�-N lines. The lowest energy �4-�1 transition is parity forbidden, and moving away from the zone center is found to result in no significant increase in the strength of optical transitions. Therefore, n-type doping will not induce strong absorption below the intrinsic optical gap, which is consistent with the high visible transparency maintained even in heavily doped In2O3. As the �8 valence band at 0.81 eV below the VBM, which determines the optical gap, exhibits almost no dis- persion and the CBM is close to parabolic, there will be a pronounced movement of the absorption onset to higher energies as the conduction band becomes more occupied; this strong shift of the optical gap is well established -1.0 0.0 1.0 2.0 3.0 4.0 E ne rg y (e V ) NH 1 4 8 (A )g (T )g g (T )u FIG. 2 (color online). Band structure of In2O3 (along the H-�-N lines). The highest energy valence bands resulting in strong optical absorption to the conduction band are light gray. A rigid shift of the conduction band is applied to offset the density- functional theory band gap underestimation [22]. 0 2 4 6 8 Energy (eV) 30 60 90 A bs or pt io n co ef fic ie nt ( 10 4 c m -1 ) Eg FIG. 3. Calculated absorption spectrum of bulk In2O3. Note that the onset of optical absorption is 0.81 eV higher in energy than the fundamental band gap [22]. PRL 100, 167402 (2008) P H Y S I C A L R E V I E W L E T T E R S week ending 25 APRIL 2008 167402-3 experimentally [1,2]. It is also worth noting that in order to preserve simultaneous transparency and conductivity, tran- sitions between the CBM band and other conduction band states near the � point must be inhibited. For In2O3, this is well obeyed as the CBM� 1 band (�5) lies 5 eV above the CBM; i.e., the transitions are well outside the visible wavelength range. We have demonstrated experimentally, using a range of x-ray spectroscopies, that the separation between the VBM and CBM is much less than the previously quoted band gap of 3.75 eV, and is not related to surface band bending; the measurements set an upper limit of 2.9 eV for the funda- mental gap. These results are supported by theoretical calculations which conclude that In2O3 is a ‘‘forbidden’’ gap material with a fundamental band gap 0.81 eV lower in energy than the onset of strong optical absorption. We therefore propose that the weak absorption observed below the optical gap of In2O3 is a property of the bulk solid, resulting from the formally forbidden transitions at the top of the valence band. Similar symmetry forbidden direct fundamental gaps have been established in other conduct- ing oxide systems including spinel SnZn2O4, SnCd2O4 and CdIn2O4 [23], CuInO2 [24], cuprite Cu2O [25], and rutile SnO2, TiO2, and GeO2 [26,27]. These results have wide consequences relating to both the basic understanding of In2O3 and its applications. For example, the recently observed unusual �1:2 eV redshift in the onset of optical absorption in nitrogen-doped In2O3 [28] (for the purposes of p-type doping) could be explained in terms of the introduction of N 2p levels above the VBM of In2O3, breaking the local symmetry and producing a parity allowed transition much higher in energy than the initial optical gap which originates from below the valence band edge. Furthermore, band offsets are generally mea- sured relative to the conduction band or valence band edges, but not both, and thus rely on prior knowledge of the band gap to determine the overall offset. Past estima- tions of In2O3 therefore need to be revised; e.g., for In2O3=Si the reported valence band offset of 1.62 eV [29] will be reduced to less than 1 eV. We are grateful to both G. W. Watson and P. J. Dobson for useful discussions. The work at Golden was supported by the U.S. Department of Energy (DOE) under Contract No. DE-AC36-99GO10337. Work in Darmstadt was sup- ported by DFG Grants No. SFB 595-D3 and No. KL1225/4 and the European Network of Excellence FAME. The Boston University (BU) program was supported in part by DOE under No. DE-FG02-98ER45680 and in part by the donors of the American Chemical Society Petroleum Research Fund. The BU XES/RIXS spectrometer system was funded by the U.S. Army Research Office under No. DAAD19-01-1-0364 and No. DAAH04-95-0014. The NSLS, Brookhaven National Laboratory, is supported by DOE under Contract No. DE-AC02-98CH10886. The Oxford program was supported by EPSRC Grant No. GR/ S94148 and the Scienta XPS facility by EPSRC Grant No. EP/E025722/1. *aron_walsh@nrel.gov [1] I. Hamberg and C. G. Granqvist, J. Appl. Phys. 60, R123 (1986). [2] C. G. Granqvist and A. Hultaker, Thin Solid Films 411, 1 (2002). [3] P. Erhart, A. Klein, R. G. Egdell, and K. Albe, Phys. Rev. B 75, 153205 (2007); S. Z. Karazhanov et al., Phys. Rev. B 76, 075129 (2007), and references therein. [4] A. Klein, Appl. Phys. Lett. 77, 2009 (2000). [5] C. McGuinness, C. B. Stagarescu, P. J. Ryan, J. E. Downes, D. Fu, K. E. Smith, and R. G. Egdell, Phys. Rev. B 68, 165104 (2003). [6] Y. Gassenbauer, R. Schafranek, A. Klein, S. Zafeiratos, M. Hävecker, A. Knop-Gericke, and R. Schlögl, Phys. Rev. B 73, 245312 (2006). [7] R. L. Weiher and R. P. Ley, J. Appl. Phys. 37, 299 (1966). [8] I. Hamberg, C. G. Granquist, K. F. Berggren, B. E. Sernelius, and L. Egström, Phys. Rev. B 30, 3240 (1984). [9] H. Köstlin, R. Jost, and W. Lems, Phys. Status Solidi A 29, 87 (1975). [10] V. Christou, M. Etchells, O. Renault, P. J. Dobson, O. V. Salata, G. Beamson, and R. G. Egdell, J. Appl. Phys. 88, 5180 (2000). [11] Y. Dou, R. G. Egdell, D. S. L. Law, N. M. Harrison, and B. G. Searle, J. Phys. Condens. Matter 10, 8447 (1998). [12] M. Sachhi et al., Phys. Rev. B 71, 155117 (2005). [13] J. Nordgren, G. Bray, S. Cramm, R. Nyholm, J. E. Rubensson, and N. Wassdahl, Rev. Sci. Instrum. 60, 1690 (1989). [14] J. H. Scofield, Lawrence Livermore National Laboratory Report No. UCRL-51326, 1973. [15] P. A. Glans, T. Learmonth, K. E. Smith, J. Guo, A. Walsh, G. W. Watson, F. Terzi, and R. G. Egdell, Phys. Rev. B 71, 235109 (2005). [16] M. Marezio, Acta Crystallogr. 20, 723 (1966). [17] P. Hohenberg and W. Kohn, Phys. Rev. 136, B864 (1964); W. Kohn and L. J. Sham, Phys. Rev. 140, A1133 (1965). [18] J. P. Perdew, K. Burke, and M. Ernzerhof, Phys. Rev. Lett. 77, 3865 (1996). [19] P. E. Blöchl, Phys. Rev. B 50, 17953 (1994). [20] G. Kresse and J. Furthmüller, Phys. Rev. B 54, 11 169 (1996). [21] B. Adolph, J. Furthmüller, and F. Bechstedt, Phys. Rev. B 63, 125108 (2001). [22] To compensate for the standard DFT band gap under- estimation, a shift was applied to the conduction band to match the calculated optical band gap with the widely quoted 3.75 eV experimental gap. [23] D. Segev and S.-H. Wei, Phys. Rev. B 71, 125129 (2005). [24] X. Nie, S.-H. Wei, and S. B. Zhang, Phys. Rev. Lett. 88, 066405 (2002). [25] J. P. Dahl and A. C. Switendick, J. Phys. Chem. Solids 27, 931 (1966). [26] D. Fröhlich, R. Kenklies, and R. Helbig, Phys. Rev. Lett. 41, 1750 (1978). [27] M. Stapelbroek and B. D. Evans, Solid State Commun. 25, 959 (1978). [28] K. R. Reyes-Gil, E. A. Reyes-Garcia, and D. Raftery, J. Phys. Chem. C 111, 14 579 (2007). [29] X. Zhang, Q. Zhang, and F. Lu, Semicond. Sci. Technol. 22, 900 (2007). PRL 100, 167402 (2008) P H Y S I C A L R E V I E W L E T T E R S week ending 25 APRIL 2008 167402-4