Processing math: 100%
J. Semicond. > 2013, Volume 34 > Issue 6 > 062002

SEMICONDUCTOR PHYSICS

First principles study on the surface-and orientation-dependent electronic structure of a WO3 nanowire

Yuxiang Qin, Deyan Hua and Xiao Li

+ Author Affiliations

 Corresponding author: Qin Yuxiang, Email:qinyuxiang@tju.edu.cn

DOI: 10.1088/1674-4926/34/6/062002

PDF

Abstract: The effects of the surface and orientation of a WO3 nanowire on the electronic structure are investigated by using first principles calculation based on density functional theory (DFT). The surface of the WO3 nanowire was terminated by a bare or hydrogenated oxygen monolayer or bare WO2 plane, and the[010]-and[001]-oriented nanowires with different sizes were introduced into the theoretical calculation to further study the dependence of electronic band structure on the wire size and orientation. The calculated results reveal that the surface structure, wire size and orientation have significant effects on the electronic band structure, bandgap, and density of states (DOS) of the WO3 nanowire. The optimized WO3 nanowire with different surface structures showed a markedly dissimilar band structure due to the different electronic states near the Fermi level, and the O-terminated[001] WO3 nanowire with hydrogenation can exhibit a reasonable indirect bandgap of 2.340 eV due to the quantum confinement effect, which is 0.257 eV wider than bulk WO3. Besides, the bandgap change is also related to the orientation-resulted surface reconstructed structure as well as wire size.

Key words: WO3 nanowiredensity functional theoryelectronic band structuredensity of states

In the family of transition metal oxides, tungsten oxide has attracted great attention due to its distinctive opto-and electrochromic properties as well as excellent gas sensing performances. It is considered a promising material for a multitude of potential applications including semiconductor gas sensors, solar-energy devices, photocatalysts, field emitters, and transparent thin-film electrochromic displays[1-3]. In recent years, various nanostructured materials such as nanowires, nanorods, nanotubes, and nanobelts, have been experimentally proved to show enhanced opto-and electrical properties due to their unique structural characteristics[4]. Considerable progress has been made in the experimental studies about preparation and properties of tungsten oxide nanostructures, especially one-dimensional nanowires[3, 5, 6].

In addition to the experiments, theoretical investigations through atomistic modeling of nanostructural materials are also necessary, so as to give intrinsic properties of sample under ideal conditions, and to propose the fundamental understanding of the physical properties behind the observation at the nanoscale[7-9]. Some theoretical results about bulk WO3 have been reported. For instance, Yakovkin et al.[10] simulated the optimized structure of the bulk WO3 (001) surface with termination of (1 × 1) O, (1 × 1) WO2, c(2 × 2) O using density functional theory (DFT) and predicted the potential role of surface relaxation in formation of crystalline nano-size clusters of WO3. Based on the hybrid functions of plane wave and localized basis sets, Wang et al.[11] conducted a systematic study about the structural and electronic properties in all phases of WO3. However, despite the few pursuable examples of the bulk WO3, to date systematical and theoretical results on WO3 nanowires are seldom reported. In particular, there is no available structural model for WO3 nanowire to be proposed, which is virtually necessary and important to theoretically investigate the optical, electrochromic and gas sensing properties.

In this work, aiming to propose a reasonable structural model of a WO3 nanowire, the dependence of the electronic properties, especially the band structure, the gap and the density of states (DOS) on the surface structure of WO3 nanowires (surface termination with WO2 surface, bare and hydrogenated oxygen monolayer) was first investigated in the framework of first-principles DFT calculations. Then the effects of wire size and orientation on the electronic bandgaps were studied. It is found that the O-terminated WO3 nanowire with hydrogenation can exhibit a reasonable indirect bandgap of 2.340 eV, which is 0.257 eV wider than bulk WO3, and the surface structure dependence, electronic properties also clearly depend on the wire size and orientation.

According to the experimental results[12-14], WO3 nanowires usually grow preferentially along the b-or c-axis, and at room temperature, the WO3 structure is monoclinic (γ-WO3 phase with space group P21/n(C52h)). In this work,[001]-and [010]-oriented WO3 nanowire with different surface structure and different size were modeled and calculated. The initial model of crystalline WO3 nanowire was constructed by selecting all the W and O atoms that fall within a virtual prism placed in bulk WO3 monoclinic crystal (room-temperature γ-WO3 phase with lattice parameters of a = 7.297 Å, b = 7.539 Å, c = 7.688 Å and β = 90.91[14]). All W and O atoms falling outside this virtual prism were removed. The direction of the prism was chosen to produce wires with [001] and [010] growth direction. Thus, termination of the WO3 nanowire surface can be accomplished by either a WO2 plane or an oxygen monolayer. Figure 1(a) and 1(b) show the top views of the initial O-and WO2-terminated WO3 nanowires with the smallest size along the [001] orientation and Figs. 1(c) and 1(d) along the [010] orientation. For the O-terminated WO3 nanowire as shown in Figs. 1(a) and 1(c), there was one dangling bond in each O atom, which can be passivated by a H atom[15]. In the present study, both bare and hydrogenated WO3 nanowires terminated by an oxygen monolayer are discussed respectively. The latter was constructed through auto-updating H atoms onto the hanging bond of the surface oxygen. When the W atoms were exposed on the surface of the nanowire, auto-updating of hydrogen was not able to realize and the nanowire was actually terminated by a "bare" WO2 surface (Figs. 1(b) and 1(d)).

Figure  1.  The top view of the [001]-(up) and [010]-oriented (bottom) WO3 nanowire terminated by (a, c) oxygen monolayer, and (b, d) WO2 plane. The dark and gray spheres represent W and O atoms, respectively.

A super cell method was employed where each wire was periodically repeated along the growth direction. To eliminate the interaction between neighboring wires, the square supercell was set to be sufficiently large with a vacuum of 10 Å in thickness[16]. In fact, it is found that a 5-Å separation between two closest facets is sufficient to make the interaction negligible[17]. The initial geometries were then relaxed to their closest. After full structural relaxation, the band structure and density of states were calculated for all the as-constructed nanowire models.

The structural optimizations, band structure, and electronic density of states were preformed by using the first principles calculations based on density functional theory (DFT). All calculations were carried out with CASTEP (Cambridge Sequential Total Energy Package) code in Materials Studio of Accelrys Inc. The ultra-soft pseudo-potential and the widely used local density approximation (LDA) with the exchange correlation functional parameterized by Ceperley and Alder (CA-PZ) were adopted. The Brilloiouin zone k-point sampling was performed using a 1 × 1 × 5 Monkhorst-Pack (MP) mesh, and the cutoff energy of plane waves is set to 340 eV. The convergence criteria for structural optimization were set to be medium quality with the tolerance for self-consistent field (SCF), energy, maximum force, and maximum displacement of 2.0 × 106 eV/atom, 2.0 × 105 eV/atom, 0.05 eV/Å and 2.0 × 103 Å, respectively.

The band structures of WO3 nanowires corresponding to the previously mentioned three surface structures shown in Figs. 2(b)-2(d) revealed their sensible dependence on the terminator, i.e., surface structure of the nanowire. The WO2-terminated nanowire (Fig. 2(b)) showed a direct bandgap (at the Q point) of 0.212 eV and larger dispersion of the conduction band (CB) than the O-terminated nanowires. Figure 2(c) indicated that the nanowire with termination of hydrogenated oxygen monolayer had an indirect bandgap, with the valence-band maximum (VBM) located at the Q point and the conduction-band minimum (CBM) located at the G point. The calculated bandgap for this kind of hydrogenated nanowire was 2.340 eV. However, the bare O-terminated nanowire (Fig. 2(d)) exhibited metallic characteristic. It can be seen that the bandgap was inconspicuous and the conduction band and valence band almost confuse.

Figure  2.  Electronic band structures of (a) bulk WO3, (b) [001]-oriented WO3 nanowires with termination of WO2 plane, (c) hydrogenated oxygen monolayer, and (d) bare oxygen monolayer.

As the dimensions of WO3 are reduced from the bulk to one-dimensional nanowire geometry, quantum confinement will result in the increase of the bandgap of the material[18]. This has been confirmed by the published experimental results. The as-prepared ultrafine tungsten oxide nanowires in the diameter range of 1-4 nm showed much larger bandgap than that in bulk form (3.36 and 2.62 eV respectively)[19, 20]. However, Figures 2(a) and 2(b) indicate that the bandgap of the WO2-terminated WO3 nanowire was markedly narrower than bulk WO3. Figures 3(a)/3(b) and 3(c)/3(d) respectively show the top view/side view of the initial and optimized [001]-oriented WO3 nanowires terminated by WO2 surface. It can be seen that, as a result of structural optimization, the O atoms on WO2 surface tend to shift outward, resulting in the W-O bond length changing from 1.895 to 1.866 Å and the W-W bond length changing from 3.705 to 2.577 Å. These changes caused the redistribution of energy states. The total DOS and partial DOS (PDOS) of O-2p and W-5d states of the WO2-terminated WO3 nanowires are shown in Fig. 4. The total DOS shown in Fig. 4(a) indicate that there existed a much high density of states at the Fermi level, which were responsible for the gap narrowing. A comparison of PDOS of O-2p and W-5d states for WO2-terminated nanowire as well as W-5d states for bulk WO3 revealed the reason of the appearance of surface states at Fermi level for the nanowire. When WO3 was terminated by WO2 surface, the W-5d states moved toward valence band (Fig. 4(b)) and resulted in the relatively high electronic states at Fermi level. That is, the surface states were mainly induced by 5d states of W atoms at WO2 surface. Therefore, it is just the "bare" WO2 surface that leads to the substantially narrow gap for a WO2-terminated WO3 nanowire.

Figure  3.  The top view (up) and side view (bottom) of (a, b) the initial and (c, d) optimized [001]-oriented WO3 nanowires terminated by WO2 surface.
Figure  4.  Total DOS and PDOS of WO2-terminated WO3 nanowire and bulk WO3 oriented along [001] direction.

Figure 5 shows the side view of the initial and optimized O-terminated [001] WO3 nanowire, and (a, b) and (c, d) respectively correspond to the hydrogenated and bare nanowire. For the WO3 nanowire terminated by hydrogenated oxygen monolayer, structural optimization resulted in a substantial relaxation of the structure. As shown in Fig. 5(b), the H atoms tended away from W atoms because of the electrostatic repulsion between H and W ions. Such a surface relaxation leads to a much stable structure. Figure 6 shows the DOS and PDOS of the WO3 nanowire terminated by hydrogenated oxygen monolayer. Differing from the WO2-terminated nanowire, the O-terminated WO3 nanowire with hydrogenation showed a bandgap of 2.340 eV, which is 0.257 eV wider than bulk WO3. This result is expected and can be well fitted to quantum confinement effect confirmed by previous experimental investigations[18-20]. Nevertheless, the experimental bandgap of 3.36 eV reported for a tungsten oxide nanowire[19] is larger than our calculated value. The reason for such a disagreement is the well-known shortcoming of the theory frame of DFT-LDA to underestimate bandgaps[21].

Figure  5.  The side view of the initial (up) and optimized (bottom) [001]-oriented WO3 nanowires terminated by (a, b) hydrogenated and (c, d) bare oxygen monolayer. The white spheres represent H atoms.
Figure  6.  (a) DOS and (b) PDOS of the [001]-oriented WO3 nanowire with termination of hydrogenated oxygen monolayer.

As seen from the DOS shown in Fig. 6(a), as the dimensions of WO3 are reduced from the bulk to a nanowire geometry, quantum confinement mainly results in the conduction band edge, which was mainly contributed by the 5d orbits of W atoms as shown in Fig. 6(b), shifts to high energy and thereby leads to the effective gap enlargement of the O-terminated nanowire. The 1s orbits of H atoms had little contribution to both conduction and valence band edges.

When the dangling bonds of the surface oxygen atoms were not passivated, a radically different band structure as shown in Fig. 2(d) was obtained. The metallic characteristic of WO3 nanowire based on this structural model can be understood from the change of electronic states in close vicinity to the Fermi level originating from the optimized nanowire structure. As it is seen from a structural comparison between the initial and optimized bare O-terminated WO3 nanowires (Figs. 5(c) and 5(d)), the optimization resulted in the formation of O=O between two adjacent slabs. The formed O=O structure can contribute a very large density of states just at the Fermi level, as illustrated in Fig. 7 which shows the PDOS of an O=O structure. Therefore, according to the calculated results described above, it is necessary to terminate the dangling bonds on the surface of the O-terminated WO3 nanowire in order to avoid the formation of O=O structure. As illustrated in Fig. 2, the WO3 nanowire with termination of hydrogenated oxygen monolayer showed a reasonable electric band structure with an indirect bandgap of 2.340 eV, which is well fit to quantum confinement effect in nanowire.

Figure  7.  PDOS of a O=O structure.

To further investigate the effects of the wire size and orientation on the electronic properties of a nanowire, the band structure of the hydrogenated oxygen monolayer-terminated WO3 nanowires with different sizes and orientations ([001] and [010] respectively) were calculated. Figure 8 shows the results obtained from [001]-oriented nanowires. Experimentally, the general diameter of as-synthesized tungsten oxide nanowire is about several nanometers and the minimum value ever reported is 1 nm[6, 19]. In this work, the bare and stoichiometric WO3 nanowire studied are respectively made up of 40, 82 and 144 atoms (W and O) in a unit cell, and the corresponding nanowire sizes are 0.9, 1.7 and 2.5 nm, respectively. Figures 8(a), 8(b) and 8(c) show the three types of nanowire structures ([001] orientation). The WO3 nanowire has undergone a transition from an indirect to direct bandgap with the increase of nanowire size, which is related to the different magnitude of the energy shift for CBM and VBM produced by quantum confinement for each point in the band structure[18]. The WO3 nanowire made up of 40 atoms has an indirect bandgap, while the bandgaps of the nanowires with 82 and 144 atoms translate to direct (at the G point).

Figure  8.  The top view and the corresponding band structures of the [001]-oriented WO3 nanowire terminated by hydrogenated oxygen monolayer with different diameters.

As shown in Figs. 2 and 8, variations in nanowire surface structure and size can affect the bandgap obviously. In addition, the bandgap of nanowire also depends on its orientation. Figure 9 further compares the calculated bandgaps of the [001]-and [010]-oriented WO3 nanowires with different sizes. As a reference, the bandgap of the bulk WO3 (2.083 eV) is also indicated by the broken line in the figure. It can be seen that the [010]-oriented WO3 nanowires showed larger bandgaps than that of bulk WO3, and the calculated bandgap increases as the size of the nanowire decreases due to the stronger quantum confinement effect in thinner wires. This observed trend shows generally good agreement with the experimental data[19, 20]. Still, the trend in the size dependence of the bandgap agrees well with those theoretical studies in Refs. [18, 22]. It has been found that the strong quantum confinement effect generally occurs when the size of the crystal is reduced to much smaller than the Bohr radius for the material (about 3 nm for WO3[23])[24]. As a result, the bandgap variation of the nanowire with size beyond 1.7 nm is no longer apparent, as shown in Fig. 9.

Figure  9.  The bandgap of WO3 nanowires with different orientations and sizes.

However, it also can be seen from Fig. 9 that, for the same wire size, the [001]-oriented nanowire shows much smaller bandgap than the [010]-oriented one, even than bulk WO3, indicating that other effects relating to the orientation of nanowire also might have remarkable contributions to the bandgap change. It has been reported that the surface effect can dominate the bandgap change of AlN nanowire and then result much smaller gap value than that of bulk solid[25]. For the [001]-oriented monoclinic WO3 nanowire, there exists an additional periodicity of alternately long and short O-O separations perpendicular to the growth direction of the nanowire, which introduced by tilting of the WO6 octahedral[26]. Structural optimization can result in the obvious change of the long and short O-O bond length and then the corresponding inclination change of the WO6 octahedral. The formed surface reconstructed structure, which is directly related to the periodicity degree of O-O separations (i.e. wire size) and the orientation of WO3 nanowire, is assumed to greatly affect the bandgap[18], although the quantum confinement effect is very important in determining the bandgap of thin nanowire (n = 40). Figure 10 shows the DOS of the [001]-and [010]-oriented nanowires with n = 40. Clearly, it is the shift of the conduction band edge to low energy as well as the shift of valence band edge to high energy that leads to the band narrowing of the [001]-oriented WO3 nanowire.

Figure  10.  DOS of the [001]-and [010]-oriented WO3 nanowires (n = 40) terminated by a hydrogenated oxygen monolayer.

The electronic structure of WO3 nanowires with various types of termination (WO2 surface, bare oxygen surface, and hydrogenated oxygen surface), different orientation, and various diameters were investigated by using the first principles calculation of plane-wave ultra-soft pseudo-potential technology based on DFT. The electronic structures of the WO3 nanowires, including their electronic density of states, band structures, and bandgaps are found to depend sensitively on the surface structure. Surface termination with hydrogenated oxygen monolayer is expected to be the most reasonable structure for a WO3 nanowire. The wire size and the orientation-resulted surface reconstructed structure can affect the bandgap of the WO3 nanowire significantly. Nanowires with smaller diameters show larger bandgaps due to a much stronger quantum confinement effect. The special surface reconstructed structure of [001]-oriented WO3 nanowire, differing from the one with [010] growth direction, can narrow the bandgap by modifying the band edges.



[1]
Gubbala S, Thangala J, Sunkara M K. Nanowire-based electrochromic devices. Sol Energy Mater Sol Cells, 2007, 91(9):813 doi: 10.1016/j.solmat.2007.01.016
[2]
Reyes L F, Hoel A. Gas sensor response of pure and activated WO3 nanoparticle films made by advanced reactive gas deposition. Sens Actuators B, 2006, 117(1):128 doi: 10.1016/j.snb.2005.11.008
[3]
Cao B, Chen J, Tang X, et al. Growth of monoclinic WO3 nanowire array for highly sensitive NO2 detection. J Mater Chem, 2009, 19:2323 doi: 10.1039/b816646c
[4]
Pan Z W, Dai Z R, Wang Z L. Nanobelts of semiconducting oxides. Science, 2001, 291:1947 doi: 10.1126/science.1058120
[5]
Baek Y, Yong K. Controlled growth and characterization of tungsten oxide nanowires using thermal evaporation of WO3 powder. J Phys Chem C, 2007, 111:1213
[6]
Sun S, Zhao Y, Xia Y, et al. Bundled tungsten oxide nanowires under thermal processing. Nanotechnology, 2008, 19:305709 doi: 10.1088/0957-4484/19/30/305709
[7]
Li Lezhong, Yang Weiqing, Ding Yingchun, et al. First principle study of the electronic structure of hafnium-doped anatase TiO2. Journal of Semiconductors, 2012, 33(1):012002 doi: 10.1088/1674-4926/33/1/012002
[8]
Gao Pan, Zhang Xuejun, Zhou Wenfang, et al. First-principle study on anatase TiO2 codoped with nitrogen and ytterbium. Journal of Semiconductors, 2010, 31(3):032001 doi: 10.1088/1674-4926/31/3/032001
[9]
Si Panpan, Su Xiyu, Hou Qinying, et al. First-principles calculation of the electronic band of ZnO doped with C. Journal of Semiconductors, 2009, 30(5):052001 doi: 10.1088/1674-4926/30/5/052001
[10]
Yakovkin I N, Gutowski M. Driving force for the WO3(001) surface relaxation. Surf Sci, 2007, 601(6):1481 doi: 10.1016/j.susc.2007.01.013
[11]
Wang F, Valentin C D, Pacchioni G. electronic and structural properties of WO3:a systematic hybrid DFT study. J Phys Chem C, 2011, 115:8345 doi: 10.1021/jp201057m
[12]
Cao B, Chen J, Tang X, et al. Growth of monoclinic WO3 nanowire array for highly sensitive NO2 detection. J Mater Chem, 2009, 19(16):2323 doi: 10.1039/b816646c
[13]
Song X, Zheng Y, Yang E, et al. Large-scale hydrothermal synthesis of WO3 nanowires in the presence of K2SO4. Mater Lett, 2007, 61:3904 doi: 10.1016/j.matlet.2006.12.055
[14]
Loopstra B O, Boldrini P. Neutron diffraction investigation of WO3. Acta Crystallogr B, 1966, 21:158 doi: 10.1107/S0365110X66002469
[15]
Gao M, You S, Wang Y. First-principles study of silicon nanowires with different surfaces. Jpn J Appl Phys, 2008, 47:3303 doi: 10.1143/JJAP.47.3303
[16]
Zhang F C, Zhang Z Y, Zhang W H, et al. First-principles study of the electronic and optical properties of ZnO nanowires. Chin Phys B, 2009, 18:2508 doi: 10.1088/1674-1056/18/6/065
[17]
Li Y, Zhou Z, Chen Y, et al. Do all wurtzite nanotubes prefer faceted ones. J Chem Phys, 2009, 130:204706 doi: 10.1063/1.3140099
[18]
Vo T, Williamson A J, Galli G. First principles simulations of the structural and electronic properties of silicon nanowires. Phys Rev B, 2006, 74(4):045116 doi: 10.1103/PhysRevB.74.045116
[19]
Hu W, Zhao Y, Liu Z, et al. Nanostructural evolution:from one-dimensional tungsten oxide nanowires to three-dimensional ferberite flowers. Chem Mater, 2008, 20:5657 doi: 10.1021/cm801369h
[20]
Gillet M, Aguir K, Lemire C, et al. The structure and electrical conductivity of vacuum-annealed WO3 thin films. Thin Solid Films, 2004, 467:239 doi: 10.1016/j.tsf.2004.04.018
[21]
Filippi C, Singh D J, Umrigar C J. All-electron local-density and generalized-gradient calculations of the structural properties of semiconductors. Phys Rev B, 1994, 50(20):14947 doi: 10.1103/PhysRevB.50.14947
[22]
Zhao X, Wei C M, Yang L, et al. Quantum confinement and electronic properties of silicon nanowires. Phys Rev Lett, 2004, 92(23):236805 doi: 10.1103/PhysRevLett.92.236805
[23]
May R A, Kondrachova L, Hahn B P, et al. Optical constants of electrodeposited mixed molybdenum-tungsten oxide films determined by variable-angle spectroscopic ellipsometry. J Phys Chem C, 2007, 111(49):18251 doi: 10.1021/jp075835b
[24]
Zheng H, Ou J Z, Strano M S, et al. Nanostructured tungsten oxide properties, synthesis, and applications. Adv Funct Mater, 2011, 21:2175 doi: 10.1002/adfm.v21.12
[25]
Zhou Z, Zhao J, Chen Y, et al. Energetics and electronic structures of AlN nanotubes/wires and their potential application as ammonia sensors. Nanotechnology, 2007, 18:424023 doi: 10.1088/0957-4484/18/42/424023
[26]
Jones F H, Rawlings K, Foord J S, et al. Superstructures and defect structures revealed by atomic-scale STM imaging of WO3 (001). Phys Rev B, 1995, 52(20):14392 doi: 10.1103/PhysRevB.52.R14392
Fig. 1.  The top view of the [001]-(up) and [010]-oriented (bottom) WO3 nanowire terminated by (a, c) oxygen monolayer, and (b, d) WO2 plane. The dark and gray spheres represent W and O atoms, respectively.

Fig. 2.  Electronic band structures of (a) bulk WO3, (b) [001]-oriented WO3 nanowires with termination of WO2 plane, (c) hydrogenated oxygen monolayer, and (d) bare oxygen monolayer.

Fig. 3.  The top view (up) and side view (bottom) of (a, b) the initial and (c, d) optimized [001]-oriented WO3 nanowires terminated by WO2 surface.

Fig. 4.  Total DOS and PDOS of WO2-terminated WO3 nanowire and bulk WO3 oriented along [001] direction.

Fig. 5.  The side view of the initial (up) and optimized (bottom) [001]-oriented WO3 nanowires terminated by (a, b) hydrogenated and (c, d) bare oxygen monolayer. The white spheres represent H atoms.

Fig. 6.  (a) DOS and (b) PDOS of the [001]-oriented WO3 nanowire with termination of hydrogenated oxygen monolayer.

Fig. 7.  PDOS of a O=O structure.

Fig. 8.  The top view and the corresponding band structures of the [001]-oriented WO3 nanowire terminated by hydrogenated oxygen monolayer with different diameters.

Fig. 9.  The bandgap of WO3 nanowires with different orientations and sizes.

Fig. 10.  DOS of the [001]-and [010]-oriented WO3 nanowires (n = 40) terminated by a hydrogenated oxygen monolayer.

[1]
Gubbala S, Thangala J, Sunkara M K. Nanowire-based electrochromic devices. Sol Energy Mater Sol Cells, 2007, 91(9):813 doi: 10.1016/j.solmat.2007.01.016
[2]
Reyes L F, Hoel A. Gas sensor response of pure and activated WO3 nanoparticle films made by advanced reactive gas deposition. Sens Actuators B, 2006, 117(1):128 doi: 10.1016/j.snb.2005.11.008
[3]
Cao B, Chen J, Tang X, et al. Growth of monoclinic WO3 nanowire array for highly sensitive NO2 detection. J Mater Chem, 2009, 19:2323 doi: 10.1039/b816646c
[4]
Pan Z W, Dai Z R, Wang Z L. Nanobelts of semiconducting oxides. Science, 2001, 291:1947 doi: 10.1126/science.1058120
[5]
Baek Y, Yong K. Controlled growth and characterization of tungsten oxide nanowires using thermal evaporation of WO3 powder. J Phys Chem C, 2007, 111:1213
[6]
Sun S, Zhao Y, Xia Y, et al. Bundled tungsten oxide nanowires under thermal processing. Nanotechnology, 2008, 19:305709 doi: 10.1088/0957-4484/19/30/305709
[7]
Li Lezhong, Yang Weiqing, Ding Yingchun, et al. First principle study of the electronic structure of hafnium-doped anatase TiO2. Journal of Semiconductors, 2012, 33(1):012002 doi: 10.1088/1674-4926/33/1/012002
[8]
Gao Pan, Zhang Xuejun, Zhou Wenfang, et al. First-principle study on anatase TiO2 codoped with nitrogen and ytterbium. Journal of Semiconductors, 2010, 31(3):032001 doi: 10.1088/1674-4926/31/3/032001
[9]
Si Panpan, Su Xiyu, Hou Qinying, et al. First-principles calculation of the electronic band of ZnO doped with C. Journal of Semiconductors, 2009, 30(5):052001 doi: 10.1088/1674-4926/30/5/052001
[10]
Yakovkin I N, Gutowski M. Driving force for the WO3(001) surface relaxation. Surf Sci, 2007, 601(6):1481 doi: 10.1016/j.susc.2007.01.013
[11]
Wang F, Valentin C D, Pacchioni G. electronic and structural properties of WO3:a systematic hybrid DFT study. J Phys Chem C, 2011, 115:8345 doi: 10.1021/jp201057m
[12]
Cao B, Chen J, Tang X, et al. Growth of monoclinic WO3 nanowire array for highly sensitive NO2 detection. J Mater Chem, 2009, 19(16):2323 doi: 10.1039/b816646c
[13]
Song X, Zheng Y, Yang E, et al. Large-scale hydrothermal synthesis of WO3 nanowires in the presence of K2SO4. Mater Lett, 2007, 61:3904 doi: 10.1016/j.matlet.2006.12.055
[14]
Loopstra B O, Boldrini P. Neutron diffraction investigation of WO3. Acta Crystallogr B, 1966, 21:158 doi: 10.1107/S0365110X66002469
[15]
Gao M, You S, Wang Y. First-principles study of silicon nanowires with different surfaces. Jpn J Appl Phys, 2008, 47:3303 doi: 10.1143/JJAP.47.3303
[16]
Zhang F C, Zhang Z Y, Zhang W H, et al. First-principles study of the electronic and optical properties of ZnO nanowires. Chin Phys B, 2009, 18:2508 doi: 10.1088/1674-1056/18/6/065
[17]
Li Y, Zhou Z, Chen Y, et al. Do all wurtzite nanotubes prefer faceted ones. J Chem Phys, 2009, 130:204706 doi: 10.1063/1.3140099
[18]
Vo T, Williamson A J, Galli G. First principles simulations of the structural and electronic properties of silicon nanowires. Phys Rev B, 2006, 74(4):045116 doi: 10.1103/PhysRevB.74.045116
[19]
Hu W, Zhao Y, Liu Z, et al. Nanostructural evolution:from one-dimensional tungsten oxide nanowires to three-dimensional ferberite flowers. Chem Mater, 2008, 20:5657 doi: 10.1021/cm801369h
[20]
Gillet M, Aguir K, Lemire C, et al. The structure and electrical conductivity of vacuum-annealed WO3 thin films. Thin Solid Films, 2004, 467:239 doi: 10.1016/j.tsf.2004.04.018
[21]
Filippi C, Singh D J, Umrigar C J. All-electron local-density and generalized-gradient calculations of the structural properties of semiconductors. Phys Rev B, 1994, 50(20):14947 doi: 10.1103/PhysRevB.50.14947
[22]
Zhao X, Wei C M, Yang L, et al. Quantum confinement and electronic properties of silicon nanowires. Phys Rev Lett, 2004, 92(23):236805 doi: 10.1103/PhysRevLett.92.236805
[23]
May R A, Kondrachova L, Hahn B P, et al. Optical constants of electrodeposited mixed molybdenum-tungsten oxide films determined by variable-angle spectroscopic ellipsometry. J Phys Chem C, 2007, 111(49):18251 doi: 10.1021/jp075835b
[24]
Zheng H, Ou J Z, Strano M S, et al. Nanostructured tungsten oxide properties, synthesis, and applications. Adv Funct Mater, 2011, 21:2175 doi: 10.1002/adfm.v21.12
[25]
Zhou Z, Zhao J, Chen Y, et al. Energetics and electronic structures of AlN nanotubes/wires and their potential application as ammonia sensors. Nanotechnology, 2007, 18:424023 doi: 10.1088/0957-4484/18/42/424023
[26]
Jones F H, Rawlings K, Foord J S, et al. Superstructures and defect structures revealed by atomic-scale STM imaging of WO3 (001). Phys Rev B, 1995, 52(20):14392 doi: 10.1103/PhysRevB.52.R14392
1

High density three dimensional integration of organic transistors

Xiaojun Guo

Journal of Semiconductors, 2019, 40(3): 030201. doi: 10.1088/1674-4926/40/3/030201

2

ZnO1-xTex and ZnO1-xSx semiconductor alloys as competent materials for opto-electronic and solar cell applications:a comparative analysis

Utsa Das, Partha P. Pal

Journal of Semiconductors, 2017, 38(8): 082001. doi: 10.1088/1674-4926/38/8/082001

3

Analysis of morphological, structural and electrical properties of annealed TiO2 nanowires deposited by GLAD technique

B. Shougaijam, R. Swain, C. Ngangbam, T.R. Lenka

Journal of Semiconductors, 2017, 38(5): 053001. doi: 10.1088/1674-4926/38/5/053001

4

The electronic and magnetic properties of wurtzite Mn:CdS, Cr:CdS and Mn:Cr:CdS: first principles calculations

Azeem Nabi, Zarmeena Akhtar, Tahir Iqbal, Atif Ali, Arshad Javid, et al.

Journal of Semiconductors, 2017, 38(7): 073001. doi: 10.1088/1674-4926/38/7/073001

5

Influences of ICP etching damages on the electronic properties of metal field plate 4H-SiC Schottky diodes

Hui Wang, Yingxi Niu, Fei Yang, Yong Cai, Zehong Zhang, et al.

Journal of Semiconductors, 2015, 36(10): 104006. doi: 10.1088/1674-4926/36/10/104006

6

Effects of interface trap density on the electrical performance of amorphous InSnZnO thin-film transistor

Yongye Liang, Kyungsoo Jang, S. Velumani, Cam Phu Thi Nguyen, Junsin Yi, et al.

Journal of Semiconductors, 2015, 36(2): 024007. doi: 10.1088/1674-4926/36/2/024007

7

Modeling of current-voltage characteristics for dual-gate amorphous silicon thin-film transistors considering deep Gaussian density-of-state distribution

Jian Qin, Ruohe Yao

Journal of Semiconductors, 2015, 36(12): 124005. doi: 10.1088/1674-4926/36/12/124005

8

Analysis of charge density and Fermi level of AlInSb/InSb single-gate high electron mobility transistor

S. Theodore Chandra, N. B. Balamurugan, M. Bhuvaneswari, N. Anbuselvan, N. Mohankumar, et al.

Journal of Semiconductors, 2015, 36(6): 064003. doi: 10.1088/1674-4926/36/6/064003

9

Electronic structures and phase transition characters of β-, P61-, P62- and δ-Si3N4 under extreme conditions: a density functional theory study

Dong Chen, Yuping Cang, Yongsong Luo

Journal of Semiconductors, 2015, 36(2): 023003. doi: 10.1088/1674-4926/36/2/023003

10

In situ TEM/SEM electronic/mechanical characterization of nano material with MEMS chip

Yuelin Wang, Tie Li, Xiao Zhang, Hongjiang Zeng, Qinhua Jin, et al.

Journal of Semiconductors, 2014, 35(8): 081001. doi: 10.1088/1674-4926/35/8/081001

11

Correlation between electrical conductivity-optical band gap energy and precursor molarities ultrasonic spray deposition of ZnO thin films

Said Benramache, Okba Belahssen, Abderrazak Guettaf, Ali Arif

Journal of Semiconductors, 2013, 34(11): 113001. doi: 10.1088/1674-4926/34/11/113001

12

A 2DEG charge density based drain current model for various Al and In molefraction mobility dependent nano-scale AlInGaN/AlN/GaN HEMT devices

Godwin Raj, Hemant Pardeshi, Sudhansu Kumar Pati, N Mohankumar, Chandan Kumar Sarkar, et al.

Journal of Semiconductors, 2013, 34(4): 044002. doi: 10.1088/1674-4926/34/4/044002

13

Modeling and simulation of centroid and inversion charge density in cylindrical surrounding gate MOSFETs including quantum effects

P. Vimala, N. B. Balamurugan

Journal of Semiconductors, 2013, 34(11): 114001. doi: 10.1088/1674-4926/34/11/114001

14

Structural parameters improvement of an integrated HBT in a cascode configuration opto-electronic mixer

Hassan Kaatuzian, Hadi Dehghan Nayeri, Masoud Ataei, Ashkan Zandi

Journal of Semiconductors, 2013, 34(9): 094001. doi: 10.1088/1674-4926/34/9/094001

15

Electronic structure and optical properties of a new type of semiconductor material:graphene monoxide

Gui Yang, Yufeng Zhang, Xunwang Yan

Journal of Semiconductors, 2013, 34(8): 083004. doi: 10.1088/1674-4926/34/8/083004

16

Optical and electrical properties of electrochemically deposited polyaniline-CeO2 hybrid nanocomposite film

Anees A. Ansari, M. A. M. Khan, M. Naziruddin Khan, Salman A. Alrokayan, M. Alhoshan, et al.

Journal of Semiconductors, 2011, 32(4): 043001. doi: 10.1088/1674-4926/32/4/043001

17

MOS Capacitance-Voltage Characteristics II. Sensitivity of Electronic Trapping at Dopant Impurity from Parameter Variations

Jie Binbin, Sah Chihtang

Journal of Semiconductors, 2011, 32(12): 121001. doi: 10.1088/1674-4926/32/12/121001

18

Neural-Network-Based Charge Density Quantum Correction of Nanoscale MOSFETs

Li Zunchao, Jiang Yaolin, Zhang Ruizhi

Chinese Journal of Semiconductors , 2006, 27(3): 438-442.

19

Electronic Structure of Semiconductor Nanocrystals

Li Jingbo, Wang Linwang, Wei Suhuai

Chinese Journal of Semiconductors , 2006, 27(2): 191-196.

20

High-Integrated-Photosensitivity Negative-Electron-Affinity GaAs Photocathodes with Multilayer Be-Doping Structures

Wang Xiaofeng, Zeng Yiping, Wang Baoqiang, Zhu Zhanping, Du Xiaoqing, et al.

Chinese Journal of Semiconductors , 2005, 26(9): 1692-1698.

  • Search

    Advanced Search >>

    GET CITATION

    Yuxiang Qin, Deyan Hua, Xiao Li. First principles study on the surface-and orientation-dependent electronic structure of a WO3 nanowire[J]. Journal of Semiconductors, 2013, 34(6): 062002. doi: 10.1088/1674-4926/34/6/062002
    Y X Qin, D Y Hua, X Li. First principles study on the surface-and orientation-dependent electronic structure of a WO3 nanowire[J]. J. Semicond., 2013, 34(6): 062002. doi: 10.1088/1674-4926/34/6/062002.
    shu

    Export: BibTex EndNote

    Article Metrics

    Article views: 2491 Times PDF downloads: 16 Times Cited by: 0 Times

    History

    Received: 07 September 2012 Revised: 20 December 2012 Online: Published: 01 June 2013

    Catalog

      Email This Article

      User name:
      Email:*请输入正确邮箱
      Code:*验证码错误
      Yuxiang Qin, Deyan Hua, Xiao Li. First principles study on the surface-and orientation-dependent electronic structure of a WO3 nanowire[J]. Journal of Semiconductors, 2013, 34(6): 062002. doi: 10.1088/1674-4926/34/6/062002 ****Y X Qin, D Y Hua, X Li. First principles study on the surface-and orientation-dependent electronic structure of a WO3 nanowire[J]. J. Semicond., 2013, 34(6): 062002. doi: 10.1088/1674-4926/34/6/062002.
      Citation:
      Yuxiang Qin, Deyan Hua, Xiao Li. First principles study on the surface-and orientation-dependent electronic structure of a WO3 nanowire[J]. Journal of Semiconductors, 2013, 34(6): 062002. doi: 10.1088/1674-4926/34/6/062002 ****
      Y X Qin, D Y Hua, X Li. First principles study on the surface-and orientation-dependent electronic structure of a WO3 nanowire[J]. J. Semicond., 2013, 34(6): 062002. doi: 10.1088/1674-4926/34/6/062002.

      First principles study on the surface-and orientation-dependent electronic structure of a WO3 nanowire

      DOI: 10.1088/1674-4926/34/6/062002
      Funds:

      the National Natural Science Foundation of China 61274074

      Project supported by the National Natural Science Foundation of China (No. 61274074) and the Tianjin Natural Science Foundation (No. 11JCZDJC15300)

      the Tianjin Natural Science Foundation 11JCZDJC15300

      More Information
      • Corresponding author: Qin Yuxiang, Email:qinyuxiang@tju.edu.cn
      • Received Date: 2012-09-07
      • Revised Date: 2012-12-20
      • Published Date: 2013-06-01

      Catalog

        /

        DownLoad:  Full-Size Img  PowerPoint
        Return
        Return