Introduction

Tungsten oxide (WO3) is widely used in industry, as catalyst and catalytic support1,2,3. Intrinsic and/or extrinsic defects can tune the compound’s behavior, in particular the electrical and optical properties, leading to electrochromic and gasochromic applications as well as to potential in areas such as smart windows, gas sensors and optoelectonics4,5,6,7. Stoichiometric WO3 is transparent and insulating with a band gap of 3.0 eV to 3.3 eV8,9, while presence of O vacancies results in optical absorption (blue color due to gap narrowing) and electrical conductivity10,11. In addition, the electronic properties, in particular the band gap, are found to be sensitive to the spatial arrangement of the W and O atoms8,9. H, due to its small size, is able to migrate in many inorganic compounds and can occupy interstitial sites without large structural expansion. It is able to induce intrinsic defects that provide free electrons12,13, modify the band gap14, interact with O vacancies15,16, and induce insulator-to-conductor transitions17. Despite much progress, H doping therefore remains a challenge to metal oxide semiconductors. On the other hand, little is known about its potential to endow semiconductors with novel electronic features.

Although the ground state of WO3 has a γ-monoclinic structure, the compound can also crystallize in other phases18. The electronic properties associated with the different structures have been investigated by density function theory, predicting that O vacancies realize a +2 charge state in the monoclinic and cubic phases19. The energy barrier for O vacancy migration turns out to be higher than 0.37 eV20. In the present work we use density functional theory to study the stability of various defects as well as their influence on the electronic structure of WO3. In addition, we report facile routes to preparing stoichiometric, O-deficient (WO3−x), and H-sufficient (HzWO3−x) tungsten oxide. We investigate the electronic and optical properties by Fourier transform infrared (FTIR) and ultraviolet-visible (UV-vis) absorption spectroscopy, combined with X-ray and ultraviolet photoelectron spectroscopy (XPS, UPS).

Results

The lattice constants of γ-monoclinic WO3 are calculated to be a = 7.27 Å, b = 7.36 Å, and c = 7.54 Å, which agrees reasonably well with the experimental values (a = 7.31 Å, b = 7.54 Å, and c = 7.69 Å) and previous theoretical results (a = 7.39 Å, b = 7.64 Å, and c = 7.75 Å)18. The structural distortions induced by defects are illustrated in Fig. 1. We observe that the O atoms surrounding a W vacancy (Vw) stay almost at their original positions, whereas nearby W atoms move towards an O vacancy (Vo), which reduces the W-W distance from 4.18 Å (perfect structure) to 3.72 Å. Serveral locations along the face and body diagonals of the WO6 unit cell are tested for possible interstitial sites. We find that a W interstitial atom is stable only at the body center (Wi) with a W-O bond length of 2.08 Å on average, which is significantly larger than in the perfect structure (1.93 Å). An O interstitial atom can be located at the body center (Oi-1) or near a W atom (Oi-2). The O-O distance of 2.12 Å in the Oi-1 case shows that there is no O-O bonding (1.21 Å in an O2 molecule), whereas in the Oi-2 case we obtain a W-O bond length of 1.92 Å. An H interstitial atom either can bond to a single O atom with a distance of 0.98 Å (Hi-1) or it can be located at the face center with O-H distances of 1.01 Å and 1.67 Å (Hi-2). Finally, an HO interstitial defect ((HO)i) is found to behave similarly to Hi-2 with O-H distances of 1.02 Å and 1.50 Å.

Figure 1: Structures: Perfect, with vacancies, and with interstitial defects, refer to the text for details.
figure 1

W, O, and H atoms are shown in black, red, and green color, respectively.

As expected, in the O-poor limit the formation energy of Vw is much higher than that of Vo, which is negative for almost all values of the Fermi level, see Fig. 2. In the O-rich limit the situation changes qualitatively only for high values of the Fermi level. We observe that Vw is neutral when the Fermi level is near the conduction band minimum, while Vo realizes a +2 charge state, in agreement with previous theoretical results19. We note that our values for the formation energy of Vo are slightly lower than those of ref. 19, which is largely due to our improved treatment of the k-mesh. For Vw we find the thermodynamic transition levels ε(0/−1), ε(−1/−2), ε(−2/−3), and ε(−3/−6) at 0.79 eV, 1.42 eV, 1.59 eV, and 1.60 eV, respectively. The formation energy of Oi-1 is higher than that of Oi-2, since the former defect does not form chemical bonds. Due to bonding with six surrounding O atoms, Wi is only stable in the +6 charge state. The formation energy in the O-poor limit is (almost) strictly negative, while Oi-1 and Oi-2 show positive values. This order changes in the O-rich limit approximately when the Fermi level exceeds the middle of the band gap. We find for both Hi-1 and Hi-2 strictly negative formation energies because of O-H bonding, reflecting easy introduction of H in WO3. Introduction of H in the form of (HO)i is possible only in O-rich environment due to the high formation energy of Oi-2. For (HO)i the thermodynamic transition levels ε(+1/0) and ε(0/−1) appear at 1.41 eV and 2.02 eV, respectively, see Fig. 2.

Figure 2: Formation energies of vacancies and interstitial defects for different charge states as functions of the Fermi level in the O-rich and O-poor limits.
figure 2

Zero Fermi level corresponds to the valence band maximum.

Figure 3 shows for perfect WO3 a (direct) band gap of 2.63 eV, in agreement with the experimental situation (2.6 eV to 3.2 eV)18 and a previous theoretical result (2.56 eV)19. The valence band maximum is almost purely due to the O 2p states and the conduction band minimum due to the W 5d states. For Vw and Oi-2 the band gap is reduced to 0.70 eV and 2.23 eV, respectively, due to the presence of in-gap states, and it becomes indirect. For Vo, Wi, and Hi-2 metallic characters are encountered, since the charge introduced by the defects enters the W 5d orbitals. Valence charge densities of the occupied (unoccupied in the case of Vw) in-gap states (entire Brillouin zone) are shown in Fig. 4. For Vw they are located on three of the six neighbouring O atoms (reflecting pronounced charge ordering), while for Vo we obtain an almost uniform distribution over all W atoms in the supercell (in agreement with ref. 19). In the cases of Wi and Oi-2, on the other hand, they are largely confined to the interstitial atoms and for Hi-2 several W atoms around the defect are involved.

Figure 3: Densities of states of perfect WO3 and of selected defective supercells, calculated with a Gaussian smearing of 0.1 eV.
figure 3

Zero energy corresponds to the valence band maximum.

Figure 4
figure 4

Charge densities of the in-gap states, see the text for details, plotted with isovalues of 0.006, 0.002, 0.006, 0.002, 0.001 electrons/bohr3 for Vw, Vo, Wi, Oi-2, and Hi-2, respectively.

Figure 5 shows W 4f XPS core level spectra obtained for films prepared in different environments. For WO3 the spectrum is deconvoluted in two peaks with weight ratio 4:3, nearly equal width of 1.7 eV, and binding energies of 36.1 ± 0.1 eV for W 4 f7/2 and 38.2 ± 0.1 eV for W 4 f5/2. The positions and shapes of these peaks agree with W in oxidation state +6, as expected for stoichiometric WO321,22. For WO3−x and HzWO3−x the W 4f signal is broadened towards lower energy, reflecting the appearance of a new oxidation state lower than +6. Deconvolution of the spectrum demonstrates two distinct contributions: The W6+ doublet as found before and a minor doublet with weight ratio 4:3, width of 1.8 eV, and lower binding energies of 34.7 eV for W 4f7/2 and 36.8 eV for W 4f5/2. The new doublet represents W+5 ions23,24, which trace back to the presence of O vacancies.

Figure 5
figure 5

Surface W 4f XPS core level spectra and their deconvolutions.

FTIR spectra, see Fig. 6(a), are measured to further elucidate the local structure changes induced by the different deposition environments. The spectra can be roughly divided into three regions: below 500 cm−1 with vibrations of the W-O bond, 500 cm−1 to 1100 cm−1 with vibrations of the W-O-W and O-W-O bonds, and above 1300 cm−1 with vibrations of the H-O-H bonds25,26. WO3 exhibits all these features with an additional shoulder near 910 cm−1, attributed to W=O bonds. The shoulder is suppressed for WO3−x and HzWO3−x, reflecting the formation of O vacancies. Peaks observed around 1600 cm−1, mainly for HzWO3−x, indicate the presence of O-H bonds, in accordance with our theoretical results. Furthermore, the UV-vis adsorption spectra in Fig. 6(b) demonstrate energy gaps of 3.0 eV, 2.8 eV, and 2.75 eV for WO3, WO3−x, and HzWO3−x, respectively.

Figure 6
figure 6

(a) FTIR and (b) UV-vis absorption spectra.

The UPS spectra in Fig. 7 show for WO3 the valence band maximum 2.9 ± 0.1 eV below the Fermi level, reflecting an n-type semiconductor, consistent with earlier experiments23. WO3−x and HzWO3−x exhibit similar valence band onsets. The main difference between the three oxides is related to the secondary electron cut-off region from which the work function can be estimated. We obtain values of 5.2 eV for WO3, WO3−x and 5.6 eV for HzWO3−x. The work function of HzWO3−x thus agrees with that of fresh tungsten oxide samples27,28, while the lower values of WO3 and WO3−x can be explained by hygroscopic water uptake that occurs instantaneously in air and reduces the work function29. Since it is believed that water interacts with O atoms in the host30,31, HzWO3−x probably maintains a high work function as the O atoms bonded with intercalated H are not available for this process. It recently has been demonstrated that annealing of WO3 in H2 atmosphere leads to O-deficient samples that show an order of magnitude enhancement in the photocurrent density32. Moreover, WO2.72 is a versatile and efficient catalyst for the hydrogenation of linear olefins, cyclic olefins, and aryl nitro groups33. In organic photovoltaics and organic light emitting diodes charge transport layers of O-deficient and H-sufficient WO3 can improve the performance of devices with forward architecture34.

Figure 7
figure 7

UPS spectra.

Intrinsic defects and H doping in WO3 have been investigated by first-principles calculations, FTIR and UV-vis absorption spectroscopy, XPS, and UPS. The favorable defect states have been established. The prediction of low Vo formation energies in O-poor environment has been confirmed by the identification of a W5+ doublet by XPS. Multiple low energy peaks in the FTIR spectrum of HzWO3−x have been attributed to vibrations of O-H bonds. UPS results on HzWO3−x have demonstrated that the work function is enhanced efficiently by H intercalation.

Methods

Density functional theory is employed based on the projector augmented wave method as implemented in the Vienna Ab-initio Simulation Package35. The generalized gradient approximation as proposed by Perdew, Burke and Ernzerhof36 (structure optimization) as well as the screened hybrid density functional proposed by Heyd, Scuseria, and Ernzerhof37 (formation energy and density of states) are used for the exchange correlation potential. The long range van der Waals interaction is taken into account by means of the DFT-D3 approach38. 2 × 2 × 2 supercells are used for all the defects to avoid artificial interaction because of the periodic boundary conditions. The cut-off energy for the plane wave basis is set to 500 eV and the energy tolerance for the iterative solution of the Kohn-Sham equations to 10−6 eV. All structures are relaxed until the residual forces on the atoms have declined to less than 0.03 eV/Å. We employ 2 × 2 × 2 k-meshes except for the hybrid density functional calculations of charged defects, for which Γ-point calculations are performed (to reduce the computational costs) and the total energy is corrected by comparison to the neutral counterparts (deviations 0.01 eV as compared to 2 × 2 × 2 k-meshes).

The defect formation energy is calculated as39

where ΔE(D, q) is the total energy difference between the perfect supercell and the supercell containing defect D in charge state q, ni is the number of atoms of type i removed from the supercell, and μi is the corresponding chemical potential. Moreover, EVBM and EF, respectively, are the valence band maximum and Fermi level (ranging from 0 eV to 2.63 eV, the size of the band gap). Stability of WO3 against byproducts and decompositions requires

The O-rich and O-poor limits are given by the maximum and minimum values of μo. Moreover, the thermodynamic transition level, for , is defined as

WO3 films are deposited in a system consisting of a stainless steel reactor with a W filament heated by an alternating current in order to vaporize its oxidized surface21. The chemical composition of the prepared oxide depends on the deposition environment: O2 (O-rich), N2 with traces of H2 (O-poor), or pure H2 (H-rich). The deposited WO3 films are characterized by FTIR absorption measurements using a Brooker spectrometer and UV-vis absorption measurements using a Perkin Elmer Lampda 40 UV/vis spectrophotometer. XPS measurements are conducted in ultra high vacuum (10−10 Torr) using a Leybold EA-11 analyzer and the unmonochromatized Mg Kα line (photon energy 1253.6 eV) at 15 keV and 20 mA anode current. The instrument is calibrated for the Au 4f7/2 peak, giving a full width at half maximum of 1.3 eV. The stoichiometry is determined from the XPS W 4f and O 1s core level spectra. After Shirley background subtraction, the photoemission peaks are integrated by fitting the O 1s and W 4f spectra with asymmetric Gaussian-Lorentzian curves. The error is estimated to be ±10%. UPS spectra are recorded for 10 nm thick films deposited on Si substrate, using the same spectrometer as for the XPS measurements and the He I excitation line (photon energy 21.22 eV). The analyzer resolution is determined to be 0.16 eV from the width of the Au Fermi edge.

Additional Information

How to cite this article: Zhu, J. et al. Intrinsic Defects and H Doping in WO3. Sci. Rep. 7, 40882; doi: 10.1038/srep40882 (2017).

Publisher's note: Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.