First-principles study on the phase transition temperature of X-doped (X = Li, Na or K) VO2

Yuanyuan Cuia, Yongxin Wanga, Bin Liua, Hongjie Luoa and Yanfeng Gao*ab
aSchool of Materials Science and Engineering, Shanghai University, Shanghai 200444, China. E-mail: yfgao@shu.edu.cn
bHuaiyin Institute of Technology, No. 1 Eastern Meicheng Rd., Huaian, Jiangsu 223003, China

Received 20th April 2016 , Accepted 26th June 2016

First published on 28th June 2016


Abstract

Vanadium dioxide (VO2) is one of the most interesting thermochromic materials with a phase transition temperature of 340 K. Our first-principles calculations indicate that Li, Na or K dopants with a doping level of 1 atomic percentage could reduce the phase transition temperature of VO2 by 43 K, 49 K, 94 K, respectively. In addition, the V–V chains feature the dimerization characteristics in the Li, Na or K doped VO2(R). The calculated electronic structures and optical properties indicate that K is an appropriate doping element for VO2, since it can effectively lower the phase transition temperature as well as enhance the near-infrared absorption.


1. Introduction

Smart windows refer to glass windows that can intelligently control the amount of light and heat passing through, usually by an external stimulus such as electrical field (electrochromic), temperature (thermochromic), ultraviolet irradiation (photochromic) and reductive or oxidizing gases (gasochromic).1 Vanadium oxides comprise an abundant and diverse family of compounds with thermochromic properties, among which V2O3, VO2 and V2O5 could undergo metal–insulator phase transitions (MIT) at the critical temperatures of 168 K, 340 K and 145 K, respectively.2,3 Table 1 summarizes the basic physical properties of V2O3, VO2 and V2O5.2
Table 1 Comparisons of the properties of V2O3, VO2 and V2O5 (ref. 2)
  Valance 3d electron Crystal structure Tc (K) Transport (T < Tc) Band gap
V2O3 V3+ 3d2 Hexagonal 168 n-type ∼0.2 eV
VO2 V4+ 3d1 Monoclinic 340 n-type ∼0.6 eV
V2O5 V5+ 3d0 Orthorhombic 145 n-type ∼2 eV


Vanadium dioxide (VO2), as a key material of thermochromic smart windows, has been consistently studied since the discovery of its MIT at ∼340 K in 1959.4,5 Across the MIT from the low-temperature monoclinic (M) phase to the high-temperature rutile (R) phase, VO2 undergoes a transition from a relative infrared transparent state to an infrared reflecting state, which is accompanied by an abrupt resistivity change.6,7 This property can be exploited to intelligently control near-infrared light radiation, which carries a large amount of solar energy, and makes VO2 a promising material for use in smart windows.8–10

The phase transition temperature (Tc) is approximately 340 K for pure VO2, and it should be lowered to near room temperature (298 K) in order to meet the requirement of the applications of VO2 for smart windows.11 To address this issue, various studies have been conducted to tailor Tc by means of element doping,12 external strain,13 photon irradiation14 or a combination of these factors.15 Reported results reveal a significant evidence that the properties of VO2 could be effectively influenced through doping special elements such as H, W, Mo or Nb.16–19 For instance, H-doped VO2 film was reported to present the metallic properties when cooled down to 120 K, and the change of Tc is around −55 K per at% H (at% means per atomic percentage).16 W is proven to be one of the most effective dopants where the transition temperature is reduced by about 28 K with increase of every 1 at% of W.17 The Mo-doped VO2 nanowires synthesized through melting-quenching method presented the decreased Tc to 315 K.18 The dopant-induced reduction of Tc in VO2 can be concisely explained as follows: the doped atoms provide or remove some of their electrons in the V 3d valence bands and make some V–V pairs break in M-phase, which leads to the destabilization of VO2 and consequently lowers the MIT temperature.

It was revealed that the reduction of Tc mentioned above is essentially attributed to injected carriers from dopants.20 More additional electrons provided from the doped atoms favor a decrease in MIT temperature. For instance, first-principles calculations indicate that Be doping can change the electronic structure by injecting additional electrons which reduce the transition temperature by 58 K per at% Be.20 Motivated by this finding, we herein propose a new strategy to modulate the transition temperature of VO2 by doping with the most frequently used alkali, i.e., Li, Na or K, which possesses one electron in the outermost orbital. Doping at an interstitial site can inject this electron into a VO2 supercell, which may reduce MIT temperature.

2. Computational details

Calculations were conducted with the Vienna ab initio simulation package (VASP), a plane wave density functional code.21,22 The potentials were of the projector augmented wave (PAW) type, and the exchange–correlation part of the density functional was treated within the generalized gradient approximation (GGA) of Perdew–Burke–Ernzerhof (PBE).23,24 To account for strong on-site Coulomb repulsion among the V 3d electrons for VO2(M),25,26 the Hubbard parameter U was added to the PBE functional in the rotationally invariant approach of Dudarev et al., in which only the difference (Ueff = UJ) between the Coulomb repulsion U and screened exchange J parameters must be specified.27,28 Despite the fact that more sophisticated methods such as GW, local density approximation with dynamical mean field theory (LDA+DMFT) and particular hybrid functionals might yield better results,29–32 the Dudarev DFT+U method has been shown to reasonably describe the electronic structure and strong correlation of VO2(M). For instance, Zhang et al. correctly predicted that both VO2(R) and VO2(M) phases were energetically stabilized by Be doping using the GGA+U method,20 and Sun et al. successfully reproduced that W, Mo, and Re were the most effective dopants to reduce the Tc of VO2.33 As discussed in references,17,20 Ueff in GGA+U formalism was chosen to be 3.4 eV in the present calculations.

The valence electron configurations for the elemental constituents are as follows: V-3d34 s2, O-2s22p4, Li-2s1, Na-3s1, K-4s1. The cut-off energy for the plane-wave basis was 520 eV. The basic calculations were performed in the supercells with 2 × 2 × 4 and 2 × 2 × 2 primitive unit cells for the VO2(R) and VO2(M) phase, respectively. Each supercell consisted of 96 atoms, with one Li, Na or K atom doped at one of three different interstitial sites: the tetrahedral (T), octahedral (O) and side centered (S) site as shown in Fig. 1. In comparison, the doped atom substituted the V site was also taken into considerations. 2 × 3 × 3 k-point meshes were used to sample the Brillouin zone for the different sized supercells. This set of parameters assured that the total energies converged to 1 × 10−5 eV per unit cell. Lattice constants and internal coordinates are fully optimized until Hellmann–Feynman forces became less than 0.01 eV Å−1.


image file: c6ra10221b-f1.tif
Fig. 1 Tetrahedral (a), octahedral (b) and side-centered (c) interstitial sites in VO2. The V, O and doped atoms are indicated by gray, red and purple spheres, respectively.

3. Results and discussion

3.1 Reduction in phase transition temperature

To compare the relative stability among the alkali dopants in the different interstitial sites of VO2, interstitial formation energy, Ei, is calculated via
 
Ei = EdopedEpurexEalkali, (1)
where Epure and Edoped correspond to the total energy of the pure and doped VO2 supercell, respectively. Ealkali is the total energy per atom of the optimized Li, Na or K metal, and x is the number of doped atoms in one supercell.

Table 2 contains the available calculated and experimental results of the lattice constants, interstitial formation energies (Ei) of Li, Na or K at different interstitial sites, supercell volumes, enthalpies and band gaps of VO2. It can be seen from the table that the enthalpies of VO2(M) and VO2(R) are −699.613 eV per supercell and −704.748 eV per supercell, respectively, which implies that the VO2(R) is more stable than VO2(M), agreeing well with the experiments.34 Moreover, Li, Na or K doping into VO2(R) or VO2(M) interstitial sites is an exothermal process due to the reason that all Ei values are negative. Then, we compare the Ei of different doping interstitial sites. In fact, we set a tight convergence criterion in the calculations which assured that the relaxation of the electronic loop stopped when the total energies change between two steps were smaller than 1 × 10−5 eV per unit supercell, and the lattice constants and internal coordinates were fully optimized until Hellmann–Feynman forces became less than 0.01 eV Å−1, so that the energies are calculated with high accuracy. Therefore, the energy differences of dopants at different sites should not be treated as errors although the values are small.

Table 2 The lattice parameters (a, b, c, α, β, γ), interstitial formation energies (Ei) of Li, Na and K in VO2 at the tetrahedral (T), octahedral (O) and side centered (S) sites, volumes (V), calculated enthalpies (H), and band gaps (Eg) collected from the experiments and the ab initio calculations
  a (Å) b (Å) c (Å) α (°) β (°) γ (°) V3 per atom) Ei (eV per atom) H (eV per supercell) Eg (eV)  
R-Phase
V32O64 9.108 9.108 11.428 90 90 90 9.872 0.000 Expt.36
V32O64 9.305 9.305 11.154 90 90 90 10.060 0.000 Calc.37
V32O64 9.313 9.313 11.156 90 90 90 10.079 −699.613 0.000 This work
LiV32O64 (T) 9.265 9.323 11.165 90 90 90 10.046 −2.351 −709.837 0.000 This work
LiV32O64 (O) 9.276 9.359 11.127 90 90 90 10.062 −2.345 −709.274 0.000 This work
LiV32O64 (S) 9.386 9.307 11.148 90 90 90 10.145 −2.345 −709.279 0.000 This work
NaV32O64 (T) 9.272 9.342 11.134 90 90 90 10.046 −2.329 −707.119 0.000 This work
NaV32O64 (O) 9.291 9.376 11.089 90 90 90 10.062 −2.325 −706.748 0.000 This work
NaV32O64 (S) 9.396 9.327 11.113 90 90 90 10.145 −2.326 −706.850 0.000 This work
KV32O64 (T) 9.274 9.265 11.224 90 90 90 10.046 −2.683 −741.220 0.000 This work
KV32O64 (O) 9.375 9.257 11.133 90 90 88.8 10.062 −2.686 −741.477 0.000 This work
KV32O64 (S) 9.289 9.402 11.154 90 90 91.3 10.145 −2.682 −741.082 0.000 This work
[thin space (1/6-em)]
M-Phase
V32O64 9.052 10.766 11.506 122.6 90 90 9.839 0.590 Expt.36
V32O64 9.100 10.600 11.480 122.2 90 90 9.760 Calc.37
V32O64 9.232 10.867 11.212 121.7 90 90 9.972 −704.748 0.610 This work
LiV32O64 (T) 9.249 10.924 11.225 121.6 90 90 10.067 −2.384 −713.030 0.734 This work
LiV32O64 (O) 9.255 10.930 11.230 121.6 90 90 10.084 −2.384 −713.026 0.738 This work
LiV32O64 (S) 9.249 10.924 11.225 121.6 90 90 10.067 −2.382 −713.030 0.715 This work
NaV32O64 (T) 9.232 10.932 11.247 121.6 90 90 10.067 −2.308 −705.141 0.487 This work
NaV32O64 (O) 9.262 10.937 11.207 121.5 90 90 10.084 −2.359 −710.036 0.714 This work
NaV32O64 (S) 9.304 10.883 11.213 121.4 90 90 10.092 −2.357 −709.911 0.682 This work
KV32O64 (T) 9.185 10.955 11.268 121.5 90 90 10.067 −2.621 −739.690 0.456 This work
KV32O64 (O) 9.335 10.876 11.188 121.5 90 90 10.084 −2.695 −742.356 0.562 This work
KV32O64 (S) 9.262 10.924 11.187 121.1 90 90 10.092 −2.683 −741.198 0.489 This work


It can be seem from Table 2 that VO2(M) with Li, Na and K at the T site, the O site and the O site has the lowest Ei, respectively, suggesting the most possible doping site of Li, Na and K in VO2 is the T site, the O site and the O site, respectively. The interstitial size of the T site and O site in VO2 are 0.70 Å and 1.21 Å, respectively. The doping Li, Na or K atom has a radius of 0.68 Å, 0.97 Å and 1.33 Å, respectively.35 The atomic radius of Li is close to that of the T site and the distortion caused by Li atom at the T site is found to be less than that at the O site, accordingly, the Li atom is easier to occupy the T site in VO2. For Na or K doped systems, the atomic radius of the dopant is larger than that of the T site but smaller than that of the O site, therefore the Na or K atom is easier to occupy the O site in VO2. For the sake of simplicity, the results of VO2 with Li, Na and K at the octahedral (T) site, the two octahedral (O) sites were taken as representatives in the following discussion, respectively.

The influence of the Li, Na or K on the phase transition temperature (Tc) can be quantitatively calculated according to the thermodynamic model of Netsianda et al.,38 in which the Tc upon doping is given via

 
image file: c6ra10221b-t1.tif(2)
where Tc,0 is the transition temperature of pure VO2, and the value of 340 K was adopted. ΔH and ΔH0 are the enthalpy changes associated with the phase transition for doped and pure VO2, respectively.38 The enthalpy was approximated as the Helmholtz free energy obtained from our DFT calculations by neglecting the pV term for condensed matter and omitting the contribution from entropy at 0 K. This thermodynamic model of Netsianda et al. has been successfully applied to a number of research results.17,20 For instance, Zhang et al. estimated the reduction of Tc to be 27 K per at% by W doping, which is quite close to 20–26 K per at% observed in experiments.17 The transition temperature change between the doped and pure VO2 can be defined as
 
ΔTc = TcTc,0 (3)

From eqn (2) and (3), it can be calculated that Li, Na or K dopants with a doping level of 1 atomic percentage could reduce Tc by 43 K, 49 K, 94 K, respectively.

3.2 Atomic and electronic structures on pure and doped VO2

The V–V chains along c-axis described in Fig. 2(a) shows that the V atoms arrange uniformly with a distance of 2.789 Å in pure VO2(R). Fig. 2(b) shows that the V–V bonds arrange alternately at 3.125 Å and 2.654 Å along c-axis due to the existence of Peierls distortion in pure VO2(M). As the three doping structures present the same change trend, we take the K-doped structure as an example. Fig. 2(g) shows the structural distortion of K-doped VO2(R) phase, the distance of V–V bonds range from 2.935 Å to 3.026 Å, and the short ones range from 2.486 Å to 2.692 Å. Fig. 2(h) describes the short V–V distance in the K-doped VO2(M) range from 2.468 Å to 2.523 Å, and the long distance range from 2.860 Å to 3.119 Å. It can be seen from Fig. 2 that all of the Li, Na or K-doped VO2(R) phases all feature the dimerization characteristics as that in VO2(M) phase. Moreover, the Li, Na or K doping has a greater influence on the atomic structures of VO2(R) with respect to that of VO2(M).
image file: c6ra10221b-f2.tif
Fig. 2 The V–V chains along c-axis for pure VO2(R) (a), pure VO2(M) (b), Li-doped VO2(R) (c), Li-doped VO2(M) (d), Na-doped VO2(R) (e), Na-doped VO2(M) (f), K-doped VO2(R) (g) and K-doped VO2(M) (h). The V, O and doped atoms are indicated by gray, red and purple spheres, respectively.

In this section, we will discuss the electronic structures on the pure and doped VO2. Fig. 3 describes the density of states (DOS) of pure VO2 and Li, Na or K doped VO2. It can be seen from Fig. 3(a), (c), (e) and (g) that the pure, Li, Na and K doped VO2(R) exhibit the metallic properties since the Fermi level crosses the conduction band. Fig. 3(b) is the density of states of the pure VO2(M), showing that the Fermi level of exists in the top of valence band with a band gap of 0.61 eV, close to the X-ray photoemission data39 and other theoretical calculation results.39,40 Fig. 3(d) and (f) show a shifting up of Fermi level to the edge of the conduction band and a widened band gap of 0.734 eV and 0.714 eV. Fig. 3(h) shows that K-doped VO2(M) has a narrowed band gap of 0.562 eV. Generally, the intrinsic band gap has an impact on the near-infrared (800–1500 nm) absorption of the VO2 film, the widened band gap for Li or Na doped VO2(M) is expected to result in an decrease in the near-infrared absorption, while the narrowed band gap for K doped VO2(M) may lead to an increase in the near-infrared absorption. Based on the discussion above, the K can be selected as an appropriate doping element for VO2, since it can lower the phase transition temperature as well as enhance the near-IR absorption.


image file: c6ra10221b-f3.tif
Fig. 3 The density of states (DOS) of pure VO2(R) (a), pure VO2(M) (b), Li-doped VO2(R) (c), Li-doped VO2(M) (d), Na-doped VO2(R) (e), Na-doped VO2(M) (f), K-doped VO2(R) (g) and K-doped VO2(M) (h). The zero point of the energy axis corresponds to the Fermi level EF.

Fig. 4 shows the V 3d partial DOS in both pure VO2 and K-doped VO2. In the atomic structure of VO2, each V atom of VO2 is surrounded by a distorted VO6 octahedron. This structure leads to a splitting of the d level to doubly degenerate eσg and triply degenerate t2g states. The two components states are labeled as dx2y2 and dz2, whereas the three components of the triply degenerate t2g states are signed as dxy, dyz, and dxz.40 Fig. 4(a) and (b) show the pure and K-doped VO2 with blue and red line, respectively. The occupied states of the doped system all extend to the lower energies, showing that the V 3d orbitals have stronger bonding. As mentioned above, the V atoms in VO2(R) are equally spaced along the c-axis, and those in VO2(M) are dimerized and tilted with respect to the c-axis (Fig. 2). We should focus on the dz2 orbital because the V–V chain extends along the c-direction which is the V–V chain direction in VO2. Fig. 4(a) shows that the occupied states of dz2 shift down to −1.26 eV and the unoccupied states raise up to 1.54 eV in K-doped VO2(R) with respect to that of the pure system, confirming a stronger bonding between the nearest-neighboring V atoms.


image file: c6ra10221b-f4.tif
Fig. 4 The partial density of state (DOS) of the V-3d orbital of pure and K-doped VO2(R) (a) and VO2(M) (b). The zero point of the energy axis corresponds to the Fermi level EF in (a) and valence band maximum EVBM in (b), respectively.

3.3 Absorption coefficients of the pure and the K-doped VO2(M)

The optical properties of a material can be described by means of the dielectric function: ε(ω) = ε1(ω) + iε2(ω).35 The imaginary part (ε2(ω)) of the dielectric function can be regarded as detailing of the real transitions between the occupied and unoccupied electronic states. The real part (ε1(ω)) of the dielectric function can be obtained from the imaginary part with the famous Kramers–Kronig relationship. The absorption coefficient (α(ω)) can be derived from ε1(ω) and ε2(ω), which are calculated via:35
 
image file: c6ra10221b-t2.tif(4)

Since our concerned VO2(R) and VO2(M) are both optically anisotropic, the components of the dielectric function, corresponding to the electric field parallel to (notated as E) and perpendicular to (notated as E) the V–V chains, have been considered in our calculations. Fig. 5 displays the absorption coefficients of the pure and the K-doped VO2(M) for the near infrared light. Fig. 5(a) shows a weak absorption for the pure VO2(M), indicating that infrared light could easily penetrate the VO2(M), in good agreement with the experimental observations. When the systems are doped with K, the VO2(M) exhibits a larger absorption coefficient for the light with energy less than 1.6 eV in Fig. 3(b), suggesting the introduction of K increases the absorption of the infrared light in VO2(M). The optical properties of the K-doped VO2(M) can be explained through its electronic structures. As displayed in Fig. 3(h), there is a 0.562 eV band-gap near the bottom of conduction band of K-doped VO2(M). The infrared light would lead to electronic transitions from the occupied state to the unoccupied states, which just corresponds to the strong absorption for the light with energy less than 1.6 eV.


image file: c6ra10221b-f5.tif
Fig. 5 The absorption spectra in the infrared spectral range obtained from first principles calculations for the pure VO2(M) (a) and K-doped VO2(M) (b). The absorption coefficient α(ω) is given in 105 cm−1.

4. Concluding remarks

In summary, by using first-principles calculation, we investigated the behavior of Li, Na or K-doping and its influence on the transition temperature (Tc) of VO2. The Li atom is easily located at the tetrahedral interstitial site in VO2, whereas the Na or K atom is easily located at the octahedral interstitial site, which can be ascribe to different atomic sizes of the dopants. The phase transition temperature of VO2 can be reduced by 43 K, 49 K or 94 K with 1% of Li, Na or K doping, respectively. The atomic structures show that the V–V chains feature the dimerization characteristics in the Li, Na or K doped VO2(R). The K can be selected as an appropriate doping alkali for VO2, since it can lower the phase transition temperature and enhance the near-infrared absorption.

Acknowledgements

The authors gratefully acknowledge the National Natural Science Foundation of China (51372228, 51325203 and 51402182), the Ministry of Science and Technology of China (2014AA032802), the Shanghai Municipal Science and Technology Commission (15XD1501700 and 14DZ2261200) and the high performance computing platform of Shanghai University.

References

  1. Y. Gao, C. Cao, L. Dai, H. Luo, M. Kanehira, Y. Ding and Z. L. Wang, Energy Environ. Sci., 2012, 5, 8708 CAS.
  2. S. Shin, S. Suga, M. Taniguchi, M. Fujisawa, H. Kanzaki, A. Fujimori, H. Daimon, Y. Ueda, K. Kosuge and S. Kachi, Phys. Rev. B: Condens. Matter Mater. Phys., 1990, 41, 4993 CrossRef CAS.
  3. T. Reeswinkel, D. Music and J. M. Schneider, J. Phys.: Condens. Matter, 2009, 21, 145404 CrossRef PubMed.
  4. F. J. Morin, Phys. Rev. Lett., 1959, 3, 34 CrossRef CAS.
  5. M. Li, D. B. Li, J. Pan, H. Wu, L. Zhong, Q. Wang and G. H. Li, J. Phys. Chem. C, 2014, 118, 16279 CAS.
  6. D. Porwal, A. C. M. Esther, I. N. Reddy, N. Sridhara, N. P. Yadav, D. Rangappa, P. Bera, C. Anandan, A. K. Sharma and A. Dey, RSC Adv., 2015, 5, 35737 RSC.
  7. M. Yang, Y. Yang, B. Hong, L. Wang, Z. Luo, X. Li, C. Kang, M. Li, H. Zong and C. Gao, RSC Adv., 2015, 5, 80122 RSC.
  8. A. I. Maaroof, D. Cho, B. Kim, H. Kim and S. Hong, J. Phys. Chem. C, 2013, 117, 19601 CAS.
  9. C. Wu, F. Feng, J. Feng, J. Dai, J. Yang and Y. Xie, J. Phys. Chem. C, 2011, 115, 791 CAS.
  10. Q. Cheng, S. Paradis, T. Bui and M. Almasri, IEEE Sens. J., 2011, 11, 167 CrossRef CAS.
  11. Y. Gao, H. Luo, Z. Zhang, L. Kang, Z. Chen, J. Du, M. Kanehira and C. Cao, Nano Energy, 2012, 1, 221 CrossRef CAS.
  12. W. Zhang, K. Wang, L. Fan, L. Liu, P. Guo, C. Zou, J. Wang, H. Qian, K. Ibrahim, W. Yan, F. Xu and Z. Wu, J. Phys. Chem. C, 2014, 118, 12837 CAS.
  13. J. Cao, E. Ertekin, V. Srinivasan, W. Fan, S. Huang, H. Zheng, J. W. L. Yim, D. R. Khanal, D. F. Ogletree, J. C. Grossmanan and J. Wu, Nat. Nanotechnol., 2009, 4, 732 CrossRef CAS PubMed.
  14. S. Lysenko, A. Rua, V. Vikhnin, F. Fernandez and H. Liu, Phys. Rev. B: Condens. Matter Mater. Phys., 2007, 76, 035104 CrossRef.
  15. S. Chang, J. B. Park, G. Lee, H. J. Kim, J. Lee, T. Bae, Y. Han, T. J. Park, Y. S. Huh and W. Hong, Nanoscale, 2014, 6, 8068 RSC.
  16. V. N. Andreev, V. M. Kapralova and V. A. Klimov, Phys. Solid State, 2007, 49, 2318 CrossRef CAS.
  17. J. Zhang, H. He, Y. Xie and B. Pan, J. Chem. Phys., 2013, 138, 114705 CrossRef PubMed.
  18. L. Q. Mai, B. Hu, T. Hu, W. Chen and E. D. Gu, J. Phys. Chem. B, 2006, 110, 19083 CrossRef CAS PubMed.
  19. C. Batista, J. Carneiro, R. M. Ribeiro and V. Teixeira, J. Nanosci. Nanotechnol., 2011, 11, 9042 CrossRef CAS PubMed.
  20. J. Zhang, H. He, Y. Xie and B. Pan, Phys. Chem. Chem. Phys., 2013, 15, 4687 RSC.
  21. G. Kresse and J. Hafner, Phys. Rev. B: Condens. Matter Mater. Phys., 1993, 47, 558 CrossRef CAS.
  22. G. Kresse and J. Furthmuller, Phys. Rev. B: Condens. Matter Mater. Phys., 1996, 54, 11169 CrossRef CAS.
  23. P. E. Blochl, Phys. Rev. B: Condens. Matter Mater. Phys., 1994, 50, 17953 CrossRef.
  24. J. P. Perdew, K. Burke and M. Ernzerhof, Phys. Rev. Lett., 1996, 77, 3865 CrossRef CAS PubMed.
  25. R. M. Wentzcovitch, W. W. Schulz and P. B. Allen, Phys. Rev. Lett., 1994, 72, 3389 CrossRef CAS PubMed.
  26. T. M. Rice, H. Launois, J. P. Pouget, R. M. Wentzcovitch, W. W. Schulz and P. B. Allen, Phys. Rev. Lett., 1994, 73, 3042 CrossRef CAS PubMed.
  27. T. J. Huffman, P. Xu, M. M. Qazilbash, E. J. Walter, H. Krakauer, J. Wei, D. H. Cobden, H. A. Bechtel, M. C. Martin, G. L. Carr and D. N. Basov, Phys. Rev. B: Condens. Matter Mater. Phys., 2013, 87, 115121 CrossRef.
  28. S. Kim, K. Kim, C. Kang and B. I. Min, Phys. Rev. B: Condens. Matter Mater. Phys., 2013, 87, 195106 CrossRef.
  29. M. S. Laad, L. Craco and E. Mueller-Hartmann, Phys. Rev. B: Condens. Matter Mater. Phys., 2006, 73, 195120 CrossRef.
  30. R. Sakuma, T. Miyake and F. Aryasetiawan, Phys. Rev. B: Condens. Matter Mater. Phys., 2008, 78, 75106 CrossRef.
  31. V. Eyert, Phys. Rev. Lett., 2011, 107, 16401 CrossRef CAS PubMed.
  32. X. Yuan, Y. Zhang, T. A. Abtew, P. Zhang and W. Zhang, Phys. Rev. B: Condens. Matter Mater. Phys., 2012, 86, 235103 CrossRef.
  33. C. Sun, L. Yan, B. Yue, H. Liu and Y. Gao, J. Mater. Chem. C, 2014, 2, 9283 RSC.
  34. G. Guzman, F. Beteille, R. Morineau and J. Livage, J. Mater. Chem., 1996, 6, 505 RSC.
  35. C. Kittel, Introduction to Solid State Physics, Wiley, New York, 2002 Search PubMed.
  36. K. D. Rogers, Powder Diffr., 1993, 8, 240 CrossRef CAS.
  37. J. Zhou, Y. Gao, X. Liu, Z. Chen, L. Dai, C. Cao, H. Luo, M. Kanahira, C. Sun and L. Yan, Phys. Chem. Chem. Phys., 2013, 15, 7505 RSC.
  38. M. Netsianda, P. E. Ngoepe, C. Richard, A. Catlow and S. M. Woodley, Chem. Mater., 2008, 20, 1764 CrossRef CAS.
  39. T. C. Koethe, Z. Hu, M. W. Haverkort, C. Schuessler-Langeheine, F. Venturini, N. B. Brookes, O. Tjernberg, W. Reichelt, H. H. Hsieh, H. J. Lin, C. T. Chen and L. H. Tjeng, Phys. Rev. Lett., 2006, 97, 116402 CrossRef CAS PubMed.
  40. A. S. Belozerov, M. A. Korotin, V. I. Anisimov and A. I. Poteryaev, Phys. Rev. B: Condens. Matter Mater. Phys., 2012, 85, 045109 CrossRef.

This journal is © The Royal Society of Chemistry 2016
Click here to see how this site uses Cookies. View our privacy policy here.