Lanli Chena,
Xiaofang Wanga,
Dongyun Wana,
Yuanyuan Cuia,
Bin Liua,
Siqi Shi*ab,
Hongjie Luoab and
Yanfeng Gao*a
aSchool of Materials Science and Engineering, Shanghai University, Shanghai 200444, China. E-mail: sqshi@shu.edu.cn; yfgao@shu.edu.cn
bMaterials Genome Institute, Shanghai University, Shanghai 200444, China
First published on 27th July 2016
Vanadium dioxide (VO2) is one of the most promising thermochromic materials with a reversible metal–insulator transition (MIT) from a high-temperature rutile phase to a low-temperature monoclinic phase, although a high MIT temperature (Tc) of 340 K for bulk VO2 restricts its wide application. Our first-principles calculations show that the oxygen nonstoichiometry plays an important role in tuning the MIT behavior of VO2. The O-vacancy in bulk VO2 gives rise to an increase in electron concentration, which induces a decrease in Tc. On the other hand, O-vacancy and O-adsorption on VO2(R) (1 1 0) and VO2(M) (0 1 1) surfaces could alter their work functions and in turn regulate Tc. In addition, the formation and adsorption energies of O-adsorption on the two types of surfaces are negative, indicating that VO2 surfaces are prone to oxidation in ambient air. The present results contribute to both tuning phase transition behaviors experimentally and reducing hindrances for the advanced applications of VO2-based materials.
Although most of the practical applications of VO2-based materials are based on its ultrafast MIT, the high phase transition temperature restricts the utilization of VO2. Many efforts have been made to tune the metal–insulator phase transition behavior of VO2, including doping,9 external strain,10 electric field,11 surface and interface engineering,12,13 and nonstoichiometry effect.14–27 Among these measures, to control oxygen nonstoichiometry of VO2 is one of the most effective routes. On the experimental side, through reactive magnetron sputtering, the resistivity ratios between the insulator and metallic phases for VO2 with the voided boundaries and oxygen vacancies are limited to 103 orders.15 Using the electrolyte gating with ionic liquids method, the electric field-induced creation of oxygen vacancies with the concentration between 0 and 50% suppresses the metal–insulator transition in VO2 films and stabilizes the metal phase to temperature below 5 K.16 Recently, Chen et al.17 found that electrolyte gating with ionic liquids on VO2 is effective to modulate the MIT behavior, which is due to the existence of oxygen-vacancy diffusion channel. Through controlling the oxygen partial pressure, the distinctive changes in the metal–insulator transition and transport behaviors were also observed.16,18 In fact, it was firstly reported in ref. 20 and 21 that the introduction of O-vacancy into VO2 reduces its Tc and even alters its electrical and optical properties. Later, Chen et al.22,23 proposed that it is the left electrons upon the formation of oxygen vacancies in VO2, corresponding to the donor doping, that cause the reduction of the phase transition temperature of VO2.28 Moreover, the high sensitivity of the MIT behavior of VO2 to O-vacancy concentration has been observed via various experimental measurements such as optical pump-terahertz probe spectroscopy, positron annihilation spectroscopy or oxide-molecular beam epitaxial method.24–26
On the theoretical side, only a few studies of oxygen-nonstoichiometry on the transition phase behavior of VO2 could be located. Using the Tran–Blaha modified Becke–Johnson (TBmBJ) generalized gradient approximation (GGA) functional, it was found that O-vacancy in VO2(M) has the higher thermodynamic stability than other defects such as V-vacancy and O-interstitial and exhibits ferromagnetism at room temperature.29 Within the generalized gradient approximation of Perdew–Burke–Ernzerhof, Mellan et al. investigated the redox behavior of the VO2(1 1 0) surface as a function of temperature and oxygen partial pressure, and found that the destabilization of VO2(R) surface should be attributed to the redox reaction occurring both at ambient conditions and at the more reducing conditions.30 The above experimental and theoretical investigations suggest that oxygen nonstoichiometry in VO2 has great impact on its phase transition behavior and phase transition temperature.
It is worth mentioning that the degrees of MIT-induced variations in work function and resistivity of VO2 are affected by oxygen nonstoichiometry31,32 and even the reduced VO2 nanobeams further induce the work function to a lower value with respect to the pure VO2.33 For instance, using Kelvin force microscopy (KFM) measurement, it is found that the work function of the VO2 surface increases from 5.15 to 5.30 eV with the increase of temperature from 300 to 368 K.32 Therefore, it is of vital importance to reveal the underlying mechanism on the effects of oxygen nonstoichiometry to the phase transition behavior for realizing the rational tuning of the Tc. In the present work, using the first-principles calculations, we study the influence of oxygen nonstoichiometry on the phase transition behavior of VO2 including phase transition temperature, electrical and optical properties, and work function, which will contribute to the general understanding of its intrinsic MIT properties and provide a theoretical reference for further experimental research.
The remainder of this paper is organized as follows. The computational details are outlined in Section 2. Section 3 details the computational models and energetics, while Section 4 discusses the influence of oxygen nonstoichiometry on the phase transition behavior of VO2 from both bulk and surface aspects. The most significant conclusions from this study are summarized in Section 5.
Valence electron configurations for the elemental constituents are as follows: V-3d34s2, O-2s22p4. The cutoff energy for the plane-wave basis is 530 eV, and Brillouin zones are sampled using 3 × 3 × 3 and 2 × 2 × 1 Monkhorst–Pack grids for bulk VO2 and VO2 surfaces, respectively. This set of parameters assures that the total energies converge to 1 × 10−5 eV per unit cell. Lattice constants and internal coordinates are fully optimized until Hellmann–Feynman forces become less than 0.01 eV Å−1. For calculation of the electronic density of states (DOS), we use a 5 × 5 × 5 k-point mesh and the linear tetrahedron method with Blöchl corrections for bulk VO2.
In general, the optical properties of the material are reflected by its transmissivity (T), the reflectivity (R) and the absorptivity (A), and their relationship can be expressed as
T + R + A = 1. | (1) |
Besides, the absorption coefficient (α(ω)) and reflectivity spectra (R(ω)) can be derived from ε1(ω) and ε2(ω), which are calculated via:
ε(ω) = ε1(ω) + iε2(ω), | (2) |
![]() | (3) |
![]() | (4) |
![]() | ||
Fig. 1 Schematic drawings of the crystal structures of (a) monoclinic and (b) rutile VO2. The V, O1 and O2 atoms are represented by gray, red and blue spheres, respectively. |
Although the various low-index rutile (1 1 0), (1 0 0), (0 0 1), (1 0 1) and (1 1 1) surfaces were observed by high-resolution TEM,45 the rutile (1 1 0) and monoclinic (0 1 1) surfaces are the energetically favorable with the lower surface energies (calculated values: 0.29 and 0.37 J m−2) compared with other surfaces.19,30 Therefore, we focus on the rutile (1 1 0) and monoclinic (0 1 1) surfaces to analyze the influence of oxygen nonstoichiometry on the phase transition behavior of VO2. Also, it is worth mentioning that the low-energy electron-diffraction shows the rutile VO2(1 1 0) surface has diffraction patterns consistent with those expected from termination of the bulk,45 i.e., no surface reconstruction was observed on the rutile VO2(1 1 0) surface. As shown in Fig. 2(a) and (c), VO2(R) (1 1 0) and VO2(M) (0 1 1) surfaces are modeled by symmetric slabs with the p(4 × 2) and p(2 × 2) unit cells, respectively. The x[0 0 1] and y[1 0] dimensions in VO2(R) (1 1 0) surface are 11.408 and 12.875 Å, respectively. While, the x[0 1
] and y[
0 0] dimensions in VO2(M) (0 1 1) surface are 11.486 and 14.002 Å, respectively. These are based on the optimized bulk lattice parameters for two phases. Our test calculations with various supercell sizes indicate that a p(4 × 2) supercell for VO2(R) (1 1 0) surface and a p(2 × 2) supercell for VO2(M) (0 1 1) ensure a formation energy of oxygen vacancy (oxygen adsorption) convergence within 0.04558 eV per f.u. (0.03251 eV per f.u.) and 0.04638 eV per f.u. (or 0.04346 eV per f.u.), respectively. Fig. 2(a) and (c) show that the slabs for VO2(R) (1 1 0) and VO2(M) (0 1 1) surfaces have 12.22 Å and 12.36 Å thickness, respectively. A vacuum layer of 18 Å is introduced to separate the slabs along the z direction. The vacuum layer guarantees that no unphysical interactions between the slab and its periodic images perpendicular to the surface influence the calculations. The bottom 5 atomic layers are fixed to simulate a bulk environment (see regions enclosed by dashed lines in Fig. 2(a) and (c)) and remaining layers are allowed to relax.
To explore the various possible adsorption models of oxygen on the VO2(R) (1 1 0) and VO2(M) (0 1 1) surface, oxygen atom is placed at the various initial sites relative to surface O or V atoms as shown in Fig. 2(b) and (d). Four adsorption sites are considered: V5c (O at the top of 5-fold coordinated V atom); V6c (O at the top of 6-fold coordinated V atom); Obr (O at the top of 2-fold coordinated O atom); O3c (O at the top of 3-fold coordinated O atom). To determine the formation of O-vacancy with different locations at the two surfaces, one oxygen atom in the first, second, and third O-atomic layer is removed, which are denoted as VObr, VO3c, and VO4c, respectively, as shown in Fig. 2(a) and (c).
The adsorption energy, Eads, of oxygen on the surfaces is calculated via
![]() | (5) |
To compare the relative stability among various configurations with O-vacancy and O-adsorption, the formation energy (Ef) is calculated via
![]() | (6) |
Table 1 summarizes the lattice parameters and band gaps of bulk VO2(R) and VO2(M). It can been seen that the calculated lattice parameters are in good agreement with the experimental values48,49 within 3% error. Our calculated band gap of pure VO2(M) (0.69 eV) is well consistent with the experimental ones of 0.6–0.7 eV (ref. 48 and 49) and the other calculation ones of 0.7–0.78 eV.50,51 Also, the lattice parameters of VO2 with O vacancy are slightly larger than those of perfect VO2. Even as the oxygen vacancy concentration increases, the calculated a, c-axis lattice parameters of VO2−x(M) (x = 0.016, 0.031) also becomes slightly larger, which is attributed to the internal compressive strains in the VO2 with O-vacancy.19,52 Fig. 3 shows the local distortion around O-vacancy (VO). As compared with pure VO2, the three V atoms around O-vacancy move outward, with the increase of V–V and V–VO distances ranging from 0.264 to 0.77 Å in VO1.984(M), and from 0.214 to 0.502 Å in VO1.984(R) (see Table 2), indicating that the oxygen vacancy induces a strong local distortion. As shown in Table 2, the short (2.519 Å) and long (3.143 Å) V–V chains arrange alternately in pure VO2(M) along a axis. Nevertheless, the short and long V–V chain lengths in VO1.984(M) range from 2.515 to 2.553 Å and from 2.781 to 3.294 Å, respectively. In VO2(R), the V–V chains (2.788 Å) arrange linearly along c axis; however, in VO1.984(R), the V–V chains are 2.501 to 2.549 Å and 2.842 to 3.354 Å, respectively, characterizing as a dimerization feature like that in VO2(M) phase, which decreases the formation energy of VO2(R) and stabilizes the VO2(R).42 Note that V1–V2 for monoclinic VO2 and V2–V3 for rutile one in Table 2 also belong to the V–V chains.
a (Å) | b (Å) | c (Å) | α (°) | β (°) | γ (°) | Eg (eV) | V (Å3) | E (eV per f.u.) | Ref. | |
---|---|---|---|---|---|---|---|---|---|---|
R-phase | ||||||||||
VO2 | 4.552 | 4.552 | 2.852 | 90 | 90 | 90 | 0 | 59.10 | — | Exp.48 |
VO2 | 4.652 | 4.652 | 2.789 | 90 | 90 | 90 | 0 | 60.34 | −21.950 | This calc. |
VO1.984 | 4.647 | 4.674 | 2.816 | 89.89 | 89.98 | 90.13 | 0 | 60.83 | −21.842 | This calc. |
VO1.969 | 4.653 | 4.653 | 2.825 | 89.99 | 90.15 | 90.07 | 0 | 61.17 | −21.717 | This calc. |
![]() |
||||||||||
M-phase | ||||||||||
VO2 | 4.517 | 5.743 | 5.375 | 90 | 122.6 | 90 | 0.60 | 117.466 | — | Exp.49 |
VO2 | 4.638 | 5.626 | 5.453 | 89.99 | 121.72 | 90.01 | 0.69 | 121.016 | −22.027 | This calc. |
VO1.984 | 4.645 | 5.626 | 5.458 | 89.57 | 121.61 | 90.06 | 0.51 | 121.453 | −21.893 | This calc. |
VO1.969 | 4.658 | 5.638 | 5.471 | 90.10 | 121.53 | 89.93 | 0.50 | 122.458 | −21.749 | This calc. |
VO2(R) | VO2(M) | VO1.984(R) | VO1.984(M) | |
---|---|---|---|---|
d1 (Å) | ||||
V1–VO | 1.937 | 1.846 | 2.151 | 2.250 |
V2–VO | 1.917 | 1.898 | 2.203 | 2.306 |
V3–VO | 1.917 | 1.994 | 2.190 | 2.258 |
V1–V3 | 3.520 | 3.661 | 3.979 | 4.212 |
V1–V2 | 3.520 | 2.524 | 3.963 | 3.294 |
V2–V3 | 2.852 | 3.532 | 3.354 | 4.122 |
![]() |
||||
dV–V (Å) | ||||
Cal. | 2.788 | 2.519 | 2.501–2.549 | 2.515–2.553 |
— | 3.143 | 2.842–3.354 | 2.781–3.294 | |
Exp.53 | 2.852 | 2.524 | — | — |
— | 3.257 | — | — |
On the other hand, it has been reported that due to a distorted VO6 octahedron surrounding each V atom in VO2, the V 3d states are split into the lower t2g (dxy, dyz, dxz) and higher eg (dz2, dx2−y2) levels.54,55 Further, given that the linear orbit clouds extend along the c-axis and the behavior of dz2 orbital determines the electronic states of rutile phase VO2,55,56 we pay more attention to the V dz2 orbital. As shown in Fig. 4(a), compared with the pure VO2(R), the occupied states of dz2 and dxy in VO1.984(R) shift down to −1.3 and −1.2 eV, respectively, and the unoccupied states of dz2 rise up to 1.22 eV, while, the states of dyz, dxz, dz2 and dx2−y2 are almost not altered. That is to say, after the introduction of O-vacancy in VO2(R), the V–V bonding states descend and the anti-bonding states ascend, which will lower the energy of VO2(R). However, in the case of VO2(M) in Fig. 4(b), only the slight difference occurs between the occupied state distributions of d orbital in VO1.984(M) and VO2(M). Therefore, the energy difference between VO2−x(R) and VO2−x(M) (x = 0.016, 0.031) is reduced due to the introduction of O-vacancy. Namely, the creation of O vacancies in VO2 is conducive to the transition from the metal phase to insulator phase.
Fig. 5(a) is adopted from ref. 20 about the band schemes of pure VO2(M). Note that the lower d∥ band is completely filled whereas the π* band is empty, and Eg is associated with their band gaps between the lower d∥ and π* band. Upon the formation of O vacancies in VO2(M) as shown in Fig. 5(b), the left electrons are located at the empty π* band, and then form the localized bands, which will lower the energy barrier and thus trigger the decrease of metal–insulator transition temperature.26,57–60 In addition, on the basis of the crystal field theory,56 the dimerization of the V atoms are paired along the c axis, which makes the V upper d∥ orbital and π* band partially occupied, and the V upper d∥ orbital overlapped by the π* band as the Fermi level (EF) moves into the conduction band, and thus induce the reduced band gap. Our calculated band gap of VO1.984(M) (0.51 eV) is lower than that of pure VO2(M) (0.69 eV), which is in good agreement with the values (0.46 eV vs. 0.60 eV) of VO2(M) with and without oxygen vacancy from the photoluminescence spectra measurement22 and other experimental results.25,58 VO2(M) with oxygen vacancy exhibits the stronger near-infrared light adsorption due to the localized nature of defect states,46 similarly to the cases of Ovac-Bi2MoO661 and C-doped TiO2.62 The same results also can be got from the density of states (DOSs) of VO2(M) and VO1.984(M) as shown in Fig. 6. For pure VO2(M) in Fig. 6(a), the conduction-band minimum is mainly contributed by the O-2p and V-3d states hybridization, while the valence-band maximum is mainly ascribed to the V-3d states mixed with a small O-2p hybridization. The Fermi level (EF) is pinned at the top of the valence band, behaving as an intrinsic characteristic. However, for VO1.984(M) as shown in Fig. 6(b), the valence and conduction bands are both shifted toward the lower energy side, and the Fermi level (EF) is located at the conduction band, presenting as n-type characteristic.
![]() | ||
Fig. 5 Outline of the band scheme of VO2 without and with O-vacancy. (a) VO2(M), (b) VO1.984(M), (c) VO2(R), and (d) VO1.984(R). |
The band schemes of VO2(R) and VO1.984(R) in Fig. 5(c) and (d) show that the π* band overlaps the d∥ band at different level and the d∥ band is occupied partially by electrons. Different from the cases of VO2(M) and VO1.984(M), the Fermi levels of VO2(R) and VO1.984(R) are transferred to the conduction band and thus their band gaps disappear.
Fig. 8(a) and (b) show the calculated and experimental imaginary parts of the dielectric function of VO2 near the infrared region, respectively. Obviously, for both phase of pure VO2, the imaginary part of dielectric function increases drastically near the infrared region. More interesting is that the imaginary part of the dielectric function of VO2 with O-vacancy has the same trend from the low temperature monoclinic phase to the high temperature rutile phase, in agreement with the experimental results of VO1.963 at 10 mTorr as shown in Fig. 8(b) from ref. 58. It is seen from the imaginary functions of pure VO2(M) in the inset of Fig. 8(a) that the main peaks occur in 1.47, 2.14, 4.84 and 6.75 eV, which are close to the experimental results pointed out its dielectric peaks between 0.25 and 5 eV.64 The reasons of optical phenomenon of pure VO2(M) are essentially stemmed from the transition between the relevant electronic states, in which the electrons at the occupied states (marked with A1) can be excited to the higher unoccupied states (marked with A2, A3, A4, and A5) in pure VO2(M) (see Fig. 6(a)). However, for VO1.984(M), as shown in Fig. 8(a), there exists a steep dielectric peak at lower energy region about 0.06 eV, which mainly derives from the transition from the occupied V-3d orbits near the Fermi level to the higher unoccupied orbits (as shown in Fig. 6(b) marked with B1, B2).
![]() | ||
Fig. 8 (a) The calculated imaginary ε2 parts of the complex dielectric functions of VO2 and VO1.984, (b) the experimental imaginary ε2 parts of the complex dielectric functions for VO2 at 10 mTorr (adapted with permission from ref. 58, Copyright 2015 Royal Society of Chemistry), (c) the absorption coefficients of VO2 and VO1.984. α(ω) is given in 105 cm−1, (d) the transmittance spectra of VO1.963 at 10 mTorr. (experimental measured: adapted with permission from ref. 58, Copyright 2015 Royal Society of Chemistry), (e) the reflectivity spectra of VO2 and VO1.984, and (f) the reflectance spectra of VO2(M) films deposited at different oxygen flow ratios (adapted with permission from ref. 65, Copyright 2014 Published by Elsevier B.V.). |
Fig. 8(c) shows the absorption coefficient of VO2 and VO1.984. It can be found that for the case with or without O-vacancy, VO2(R) shows the stronger absorption than VO2(M) near the infrared region. Being comparable with the pure rutile (or monoclinic) phase, VO1.984 exhibits a peak at 0.35 eV (or 0.23 eV) with an absorption coefficient of about 1.68 × 104 cm−1 (or 1.9 × 104 cm−1). These results also have been derived from the PDOS in Fig. 6(b), in which the existence of strong inner-band and inner-band absorption will induce the low transmittance.
The calculated and experimental65 reflectivity spectra of VO2 are shown in Fig. 8(e) and (f). It is clearly seen that the calculated values of VO1.984(M) (0.771) and VO1.984(R) (0.784) are higher than those of the pure VO2 (0.368 and 0.620), which agrees well with the experiment results. This suggests that the introduction of O-vacancy in VO2 can improve the reflectivity near the infrared region. Overall, the pure VO2(M) shows very weak absorptivity and reflectivity near the infrared region whereas the relative strong ones are true for the pure VO2(R). These aspects vividly indicate that the infrared light may easily penetrate the VO2(M), but be significantly blocked by the VO2(R). A similar conclusion was drawn by Zhang et al. who investigated the transmittance of VO2 by using the HSE06 exchange–correlation functional.41 It is worth emphasizing that although VO1.984(R) presents a larger reflectivity in the infrared range than VO2(R), VO1.984(M) exhibits large reflectivity in the light with energy less than 0.43 eV. In other words, the introduction of O-vacancy in VO2 lowers its transmissivity in the infrared range as shown in Fig. 8(d), which in turn leads to the declination of the desired optical switching to some extent.
Eads (eV) | Ef (eV) | dO–O (Å) | dV–O (Å) | dad (Å) | d12 (Å) | μ (D) | |
---|---|---|---|---|---|---|---|
a dO–O is the distance between the O adatom and its nearest neighboring O atom. dad is the distance from the O adatom to the O atom in the first layer. d12 is the thickness of the surface bilayer. The surface dipole moment μ (D) of one and two O atoms adsorption at different adsorption sites on the VO2(R) (1 1 0). | |||||||
Clean | — | — | — | 1.807–2.034 | — | 1.038 | — |
Obr | −1.373 | 2.014 | 1.313 | — | 1.313 | 1.300 | 0.982 |
V5c | −5.585 | −2.198 | — | 1.606 | 1.606 | 1.241 | 0.998 |
V6c | −5.270 | −1.883 | — | 1.607 | 1.607 | 0.659 | 1.029 |
O3c | −3.668 | −0.281 | 1.472 | — | 1.281 | 1.101 | 0.156 |
C1 | −5.276 | −3.784 | 3.078 | 1.601–1.602 | 1.600 | 1.261 | 0.725 |
C2 | −4.457 | −2.148 | 1.396 | 1.850–1.859 | 1.720 | 1.195 | 0.320 |
C3 | −4.315 | −1.862 | 1.384 | 1.934–1.896 | 1.871 | 1.113 | 0.444 |
C4 | −4.285 | −1.802 | 1.272 | 1.765–2.017 | 1.755 | 1.214 | 0.437 |
VObr | — | 3.195 | — | — | — | 1.080 | — |
VO3c | — | 2.787 | — | — | — | 0.729 | — |
VO4c | — | 2.405 | — | — | — | 0.577 | — |
Eads (eV) | Ef (eV) | dO–O (Å) | dV–O (Å) | dad (Å) | d12 (Å) | μ (D) | |
---|---|---|---|---|---|---|---|
a dO–O is the distance between the O adatom and its nearest neighboring O atom. dad is the distance from the O adatom to the O atom in the first layer. d12 is the thickness of the surface bilayer. The surface dipole moment μ (D) of one and two O atoms adsorption at different adsorption sites on the VO2(M) (0 1 1). | |||||||
Clean | — | — | — | 1.712–2.225 | — | 1.010 | — |
Obr | −1.824 | 1.563 | 1.308 | — | 1.301 | 1.302 | 0.156 |
V5c | −5.279 | −1.892 | — | 1.604 | 1.604 | 1.266 | 0.998 |
V6c | −2.065 | 1.322 | — | 1.872 | 1.458 | 1.342 | 0.374 |
O3c | −2.372 | 1.013 | 1.471 | 2.104–2.0444 | 1.276 | 1.197 | 0.203 |
C1 | −4.925 | −3.083 | 2.988 | 1.736–2.214 | 1.603–1.606 | 1.240 | 0.990 |
C2 | −3.957 | −1.146 | 1.398 | 1.706–2.231 | 1.718 | 1.215 | 0.445 |
C3 | −4.031 | −1.294 | 1.383 | 1.704–2.230 | 1.880 | 1.212 | 0.741 |
C4 | −3.750 | −0.733 | 1.264 | 1.845–2.216 | 1.730 | 1.257 | 0.593 |
VObr | — | 3.596 | — | — | — | 1.198 | — |
VO3c | — | 3.027 | — | — | — | 0.730 | — |
VO4c | — | 2.568 | — | — | — | 0.506 | — |
For O-adsorption on the VO2(R) (1 1 0) surface, the formation energy of the most stable configuration with one O-adsorption (V5c: −2.198 eV) is higher than that of the more energetically favorable configuration about two O-adsorption (C1: −3.784 eV), meaning that VO2 is prone to oxidation at the ambient condition. This will be in detail discussed in Section 4.2.2. Note that the similar results can be obtained from the O-adsorption on the VO2(M) (0 1 1) surface, which was also proved by Mellan et al.30 who reported that the rutile (1 1 0) surface corresponds to the monoclinic (0 1 1) surface.
Starting from one O atom adsorption at V5c, we examine four configurations involving two oxygen atoms adsorption on VO2(R) (1 1 0) and VO2(M) (0 1 1), which are labeled as C1, C2, C3 and C4, respectively. Their relaxed structures and adsorption energies are summarized in Fig. 9. For the VO2(R) (1 1 0) and VO2(M) (0 1 1) surfaces with two O atoms adsorption, the adsorption energy sequences follows C1 < C2 < C3 < C4 and C1 < C3 < C2 < C4, respectively. Note that the difference between C2 and C3 is slightly smaller for two surfaces (0.142 and 0.074 eV) and thus the adsorption behaviors of two O-atoms on the VO2(M) (0 1 1) and the VO2(R) (1 1 0) surfaces are similar to each other. The configuration C1 with two O atoms adsorption at the V5c is considered to be energetically more favorable, which can be explained from the relatively strong interaction between O adatoms and the surface atoms. The configuration C1 exhibits the shorter V–O bond lengths (VO2(R) (1 1 0): 1.601 and 1.602 Å; VO2(M) (0 1 1): 1.603 and 1.606 Å), which agree well with the value of 1.61 Å using PBE method30 and the values reported elsewhere for surface vanadyl species (1.55–1.70 Å).67 As shown in Fig. 9, the O–O bond lengths (VO2(R) (1 1 0): 1.272 Å; VO2(M) (0 1 1): 1.264 Å) in C1 are consistent with that of the typical O–O double bond (1.21 Å), whereas the calculated V–O bond lengths in the outermost layer (VO2(R) (1 1 0): 1.755 Å; VO2(M) (0 1 1): 1.730 Å) are both shorter than the V–O distance in the bulk (1.954 Å). Given the above characteristics about oxygen adsorption on the surfaces, there is another matter which we need to pay attention to in the actual application: VO2 exposed to ambient condition is prone to oxidation, which will degrade its thermochromic properties. Experimentally, using atomic layer deposition (ALD) method, Wang et al. deposited an ultrathin Al2O3 film with 15 nm to protect the underlying VO2 from degradation.68
Φ = x − EF, | (7) |
![]() | ||
Fig. 10 Comparison of the work functions between O-vacancy and O-adsorption on the two types of surfaces. (a) VO2(R) (1 1 0) and (b) VO2(M) (0 1 1) surface. |
However, it can be seen from Fig. 10 that the average work functions of each surface with the O-adsorption on VO2(R) (1 1 0) and VO2(M) (0 1 1) surfaces are increased by 0.20% and 0.9% for one or two oxygen adsorption higher than that of the corresponding clean surface, respectively. Consequently, as compared with the clean surface, the surface with O adsorption should exhibit the higher phase transition temperature. This can be explained from the smaller electronegativity of surface-V atom (1.63) as compared with O adatom (3.44) and the tendency that a number of electrons from V atoms on the surface layers transfer to the O adatoms.
By referring to ref. 70, the variation of the work function caused by the O vacancy (or O adsorption) is also related to the surface dipole moment via
μ = AΔΦ/(12πθ), | (8) |
Fig. 11 shows the plane-averaged electronic charge density and electron density difference along the z axis. The charge redistribution caused by oxygen adsorption can be calculated by
Δρ = ρtot − ρclean − ρO, | (9) |
Based on the above results and discussion, it is found that oxygen nonstoichiometry in VO2 can modify its phase transition behavior, which agree well with the experimental results.16–22,32,33,57 However, in the ambient air, VO2 is prone to oxidation,19,30,68 which exhibits the lower adsorption energy, indicating a strong interaction between O and the surface V-atoms. Therefore, for VO2, in actual use, the oxygen pressure should be controlled and a VO2 film should avoid being exposed to ambient air.
For O-vacancy and O-adsorption on the VO2 surfaces, the energetically favorable O-vacancy locates in the inner layer, whereas the adsorbated-oxygen atom is found to be most favorable at V5c site. Furthermore, the negative formation and adsorption energies after O-adsorption mean that VO2 surfaces can be easily oxidized. The work function of the presence of O-vacancy in the surface is lower than that of O-adsorption on the same surface, which has a great impact on the phase transition behavior of VO2. These findings pave favorable explanations about the MIT behavior of O-vacancy and O-adsorption of VO2 for specific applications such as ultrafast electronic switches and electron-optical devices.
This journal is © The Royal Society of Chemistry 2016 |