Intrinsic defects in gallium sulfide monolayer: a first-principles study

Hui Chen, Yan Li*, Le Huang and Jingbo Li*
State Key Laboratory for Superlattice and Microstructure, Institute of Semiconductor, Chinese Academy of Science, Beijing 100083, China. E-mail: bclly2008@semi.ac.cn; jbli@semi.ac.cn

Received 6th May 2015 , Accepted 26th May 2015

First published on 26th May 2015


Abstract

The electronic and magnetic properties of native point defects, including vacancies (VGa and VS), antisites (GaS and SGa) and interstitials (Gai and Si) in monolayer and bulk GaS, were systemically studied using the density functional theory method. For the monolayer, the impurity states appeared in the band gaps of all defect structures except interstitial Si. Half-metallic behavior can be obtained in the presence of VGa and Gai. Monolayers with VGa, GaS, SGa and Gai had a total magnetic moment of 1.0 μB, as did the bulk samples with VGa, GaS and SGa, whereas the monolayers with VS and Si and bulk sample with Gai were spin-unpolarized. In addition, n- and p-type GaS monolayers were obtained under Ga-rich and S-rich conditions, respectively. GaS and SGa were identified as suitable n- and p-type defects, respectively.


I. Introduction

Two-dimensional (2D) materials have attracted increasing interest as a result of their exotic transport physics and remarkable electrical, optical, magnetic and mechanical properties.1–4 The basic photovoltaic and photoelectronic properties of monolayer GaS, a III–VI semiconductor, have been extensively studied in terms of its potential application in solar cells, solid-state batteries and memory devices.5–13 GaS and GaSe monolayer field-effect transistors have been made with n- and p-type conducting features, respectively.13 A few-layer GaS two-terminal photodetector with a fast and stable response was fabricated by Yang et al.14 A higher photo-response and external quantum efficiency were obtained in NH3 than in air or O2 as a result of the opposite roles that NH3 and O2 play during charge transfer between the adsorbed gas molecules and the photodetector. Doped few-layer GaS has also been considered as a promising material for the fabrication of near-blue light-emitting devices by controlling the defects and their electronic properties during preparation.15

From the theoretical point of view, the structural and electronic properties of GaS layered compounds have been extensively studied using the density functional theory (DFT) method.16–23 Zólyomi et al.16 showed that 2D crystals of Ga2X2 (X = S, Se and Te) were dynamically stable indirect band gap semiconductors with a sombrero dispersion of holes near the valence band maximum (VBM). Ma et al.17 reported that the band gaps of GaS and GaSe increased from multilayer to single layer structures and could be tuned under mechanical strain, which makes them potential candidates for tunable nanodevices. Orudzhev and Kasumova18 established that the elastic constants of GaS layered compounds increase monotonically in the pressure range of 0–20 GPa as the pressure increases. Zhou et al.19 demonstrated that GaS nanoribbons can display an intrinsic half-metallic character with ferromagnetic coupling, raising from the Ga-4s, Ga-4p, and S-3p states. Ding and Wang20 reported characteristic Dirac-like band features in linear dispersions of Si/GaS heterosheets. They also proposed that a sizable band gap at the Dirac point is opened as a result of the intrinsic electronic field and this can be tuned by altering the voltage or strain. Understanding basic information about native and exotic defects is of great importance as structural imperfections, such as lattice defects, may change the electronic, magnetic and optical properties of materials. However, to the best of our knowledge, little work has been reported on the electronic and magnetic properties of native point defects in GaS monolayer.

We investigated the electronic and magnetic properties of intrinsic defects (VGa, VS, GaS, SGa, Gai and Si) in GaS monolayer by a first-principles method based on DFT. The formation energies under two different preparation conditions were taken into account: (i) gallium-rich and sulfur-poor conditions; and (ii) gallium-poor and sulfur-rich conditions. We found that, without any doping impurities, GaS monolayer will tend to form an n-type semiconductor under gallium-rich and sulfur-poor conditions, whereas they will tend to form a p-type semiconductor under gallium-poor and sulfur-rich conditions. By analogy, the same results are expected to be applicable to other III–VI semiconductors such as GaSe and GaTe.

II. Computational method

Calculations were performed within DFT using generalized-gradient approximation with the functional of Perdew, Burke and Ernzerhof (PBE) method24,25 implemented in the Vienna ab initio simulation package.26,27 For comparison, some calculations were also performed using the Heyd–Scuseria–Ernzerhof (HSE06) hybrid functional.28,29 The cutoff energy was set to 364 eV. For the GaS monolayer, a unit cell including four atoms was used in the calculations for the pristine GaS monolayer and a large 4 × 4 supercell with one defect was chosen for the defective monolayer. Brillouin zone sampling was performed with Monkhorst–Pack special k-point meshes.30 We chose k-point grids of 20 × 20 × 1 and 5 × 5 × 1 to calculate the pristine and defective systems, respectively. The distance between adjacent monolayers was >12 Å and hence the interaction between the monolayers could be eliminated. For bulk GaS, a unit cell containing eight atoms was chosen for the calculations of pristine bulk GaS and a 3 × 3 × 1 supercell including one defect was used for the defect system. 15 × 15 × 3 and 5 × 5 × 1 k-point grids were used in the pristine and defect bulk samples, respectively. The distance between two layers was 7.5 Å. The convergence criteria used for the electronic self-consistent relaxation and the ionic relaxation were set to 10−6 eV for energy and 0.02 eV Å−1 for force, respectively.

Typical intrinsic defects in the GaS monolayer and bulk sample, including vacancies (VGa and VS), antisites (GaS and SGa) and interstitials (Gai and Si) were studied. To search the most stable configuration of the interstitials, four different positions labeled as TS (at the top of S atoms), TG (at the top of Ga atoms), C (above the central atom of the hexagonal ring of GaS) and M (above the middle of the Ga–S bond) were considered (Fig. 1).31


image file: c5ra08329j-f1.tif
Fig. 1 Schematic structure of the GaS monolayer. The blue, yellow and green (or red) balls represent the Ga, S and intrinsic defect atoms, respectively.

To determine the formation energy and transition energy levels of a defect, the total energy ED,q of the GaS monolayer containing the defect atom α with charge state q was calculated. The formation energy ΔHD,q of the defective system was defined as:32–38

 
image file: c5ra08329j-t1.tif(1)
where EH is the total energy of the GaS monolayer without defects,32,33 nα is the number of defect atoms (nα = −1 when an atom is added and nα = 1 when an atom is removed), μα is the atomic chemical potential34–36 and q is the number of electrons transferred from the supercell to the reservoirs in the formation of the defective cell.39 EV is the energy at the VBM of the defect-free system. Here, the electron Fermi energy EF is limited between the VBM and the conduction band minimum (CBM). Consequently, the defect formation energy strongly depends on μα and EF.

μα can be also expressed as μα = μsolidα + Δμα,37,40 where μsolidα is the total energy of solid Ga or S in its stable phase. However, there are some thermodynamic limits on the achievable values of the chemical potentials under thermal equilibrium growth conditions. First, Δμα is bound by Δμα ≤ 0 to avoid the precipitation of Ga and S, corresponding to the “α-rich condition”. Second, Δμα is limited to values that maintain the stable GaS monolayer, through which an “α-poor condition” can be obtained.39

The transition level ε(D, q/q′) is the Fermi level position at which the formation energy of defect D at charge state q is equal to that at charge state q′, namely, ΔHD,q = ΔHD,q.37,41–43 From eqn (1), the transition energy level can be obtained by:

 
ε(D,q/q′) = [ΔHD,q(EF = 0) − ΔHD,q(EF = 0)]/(q′ − q) (2)
where ΔHD,q (EF = 0) is the formation energy of defect D with the charge state q when EF = 0.41,44 To avoid the problem that using special k-points or equivalent k-points in the superstructure gives a poor description of the symmetry and energy levels of the defect state, a hybrid scheme combining the advantages of both special k-points and Γ-point only approaches was used.39,45 Core level shifts were also considered.

III. Results and discussion

The hexagonal structure of GaS has a layered S–Ga–Ga–S repeating unit built by six-membered Ga3S3 rings.13 The calculated heat of formation was −1.11 eV. The pristine GaS monolayer and bulk sample were both non-magnetic indirect band gap semiconductors. The gap values of the monolayer and bulk sample were 2.58 and 1.60 eV calculated by the PBE method, which were in good agreement with other theoretical results17 and close to the experimental values of 3.05 (ref. 46) and 2.53 eV,47 respectively. For comparison, the band gaps of the monolayer and bulk sample were 3.28 and 2.30 eV, respectively, using the HSE06 method. The band structure and density of states of the GaS monolayer and bulk sample calculated by both the PBE and HSE06 methods are plotted in Fig. 2. The VBM and CBM of the GaS monolayer were located near the Γ point and at the M point, respectively, whereas in bulk GaS they were situated at the Γ and M points, respectively. In the GaS monolayer, the lower conduction bands were mainly attributed to the Ga-4s and S-3p states, whereas the lowest conduction bands were attributed to the hybridization of the Ga-4p and S-3p states (Fig. 2b). The uppermost valence bands were primarily from the Ga-4p and S-3p states in addition to a small contribution from the Ga-4s state. The valence bands in the range −3.0 to −1.5 eV were mainly contributed by the Ga-4p and S-3p orbitals, except for a small contribution from the Ga-3d states. The hybridization of the Ga-4p and S-3p orbitals were located at a deeper energy range of about −3.1 eV. The orbital components of the energy bands in bulk GaS were similar to those of energy bands in the GaS monolayer. By comparing Fig. 2a with c, and b with d, we found that the band structures and density of states of both the GaS monolayer and bulk sample calculated by the PBE method had a similar alignment to the results calculated by the HSE06 method. In view of the huge cost of computing resources and time, the standard PBE functional was considered synthetically in the calculations of the defective systems.
image file: c5ra08329j-f2.tif
Fig. 2 (a and c) Band structure and (b and d) density of states (DOS) of GaS monolayer compared with the bulk phase using the PBE and HSE06 functionals, respectively. The solid and dotted lines represent the GaS monolayer and bulk phase, respectively. EF is plotted with respect to the VBM and set to zero.

The electronic properties of GaS monolayer with typical native defects were studied. Fig. 3 shows the total density of states and projected density of states of the intrinsic defects in the GaS monolayer. This shows that VGa and Gai doped GaS monolayer are p- and n-type semiconductors, respectively. The impurity states appear in the band gaps of all the defect structures except Si; the impurity states induced by VS, GaS and SGa belong to deep-level states in the gaps. Therefore the band gap of the pristine monolayer is evidently tuned, which significantly affects its optical and transport properties. It is clear that the impurity states of VS are mainly formed by the 4p states of the Ga atom closest to the S vacancy, whereas these impurity states of GaS and SGa are attributed to hybridization between the p states of the substituted atoms and the Ga (S) atom nearest them, with a small contribution from the p states of the nearest S (Ga) atom. It can be seen that VGa and Gai behave as half-metals and can be used as spin filters. Their band structures are shown in Fig. 4. VGa in sheets behaves as an indirect p-type semiconductor for up-spin channels and as a metal for down-spin channels; the corresponding band gap is 2.18 eV from the M to the Γ point for up-spin electrons. In contrast, for Gai, the sheet is metallic for the up-spin channel and an n-type semiconductor with an indirect band gap of 2.47 eV from the k point to the Γ point for the down-spin channel.


image file: c5ra08329j-f3.tif
Fig. 3 Total density of states (TDOS) and projected density of states (PDOS) of VGa, VS, GaS, SGa, Gai and Si in GaS monolayer. The uppermost plane is TDOS (a). The planes (b), (c) and (d) are the PDOS of the Ga atom, the S atom closest to the defect position and the defect atom, respectively. The red, blue and green lines represent the s, p and d orbitals, respectively. EF is set to zero.

image file: c5ra08329j-f4.tif
Fig. 4 Band structures of VGa and Gai in GaS monolayer. The red (blue) color corresponds to the up-spin (down-spin) channel. EF is set to zero.

Bader analysis48 was used to calculate the charge transfer. For perfect bulk GaS, the loss of electrons on four Ga atoms was 3.05e. For the pristine GaS monolayer, the depletion of electrons on two Ga atoms in total was 1.52e. Table 1 shows that the antisite atoms exchanged smaller amounts of excess electrons with the monolayer compared with other types of defect atoms, especially for substituted Ga atoms in GaS; here, 0.114e was subtracted from the Ga substituted atom. The S adatom in Si exchanged the most excess electrons among all the different types of defect atoms – that is, 0.939e transferred from the monolayer onto the S adatom.

Table 1 Calculated values for GaS monolayer and bulk sample with native VGa, VS, GaS, SGa, Gai and Si defectsa
Species Features VGa VS GaS SGa Gai Si
a Δρ (e), excess charge on dopant or the nearest atom close to the vacancy; ΔE (eV), energy difference between magnetic and non-magnetic states; μtotal (μB), magnetic moments of the whole supercell and dopant or the nearest atom close to the vacancy μ (μB).
Monolayer Δρ 0.645 −0.496 −0.114 0.337 −0.543 0.939
ΔE −0.059 −0.057 −0.099 −0.001
μtotal 1.00 1.00 1.00 1.00
μ 0.19 0.24 0.14 0.10
Bulk Δρ 0.733 −0.490 −0.138 0.351 −0.685 0.296
ΔE −0.002 −0.005 −0.006
μtotal 1.00 1.00 1.00
μ 0.01 0.253 0.055


These native defects not only affect the electronic properties, but also bring about magnetic features. To determine the ground state of the GaS monolayer and bulk sample with native defects, we calculated the energy difference between the spin-polarized and spin-unpolarized states (ΔE)49 (Table 1). VGa, GaS, SGa and Gai in the GaS monolayer had a magnetic ground state, whereas VS and Si were spin-unpolarized. All the magnetic moments were 1.0 μB per supercell in VGa, GaS, SGa and Gai. However, the contributions of the nearest S atom close to the Ga vacancy in VGa, the Ga substitution in GaS, the S substitution in SGa and the Ga adatom in Gai to the total magnetic moments were unequal at 16.0, 19.4, 12.3 and 9.1%, respectively. Fig. 5 shows that the spin polarization in the monolayer was mainly localized on the dopants and neighboring atoms for the SGa antisite, whereas larger spatial extensions of spin density were found for the VGa vacancy, the GaS antisite and the Gai interstitial. In bulk GaS, VGa, GaS and SGa had a magnetic ground state with magnetic moments of 1.0 μB per supercell, whereas VS, Gai and Si were spin-unpolarized. The contributions of the nearest S atom close to the Ga vacancy in VGa, Ga substitution in GaS and S substitution in SGa to the total magnetic moments were 1.0, 20.2 and 5.2%, respectively. The reason for the Gai in the monolayer having magnetic properties while it is spin-unpolarized in the bulk sample was thought to be due to the weak Van der Waals forces between adjacent layers.


image file: c5ra08329j-f5.tif
Fig. 5 Spin density isosurfaces of GaS monolayer with native defects: (a) VGa; (b) GaS; (c) SGa; and (d) Gai. The isosurface value is set to 0.0002e Å−3. The pink and green colors represent positive and negative values, respectively.

We now discuss the formation energy of the various native defects in GaS monolayer. The vacancy VGa is taken as an example to illustrate the relationship between the formation energy and EF. Fig. 6 shows the formation energies of the acceptor point defect VGa with the charge state q of 0, 1− and 2−. As the formation energy of VGa with q = 2− is the lowest among all the charge states, it will be prone to form VGa (q = 2−). As the concentration of cavities increases, EF will decrease. During the process in which EF moves downward from the CBM to the VBM, the charge state of VGa will transfer from 2− to 1− and, finally, to 0. Table 2 lists the calculated formation energies of the intrinsic point defects with each charge state. For the neutral charge state, the formation energy of SGa is the highest, whereas that of VS is the lowest under gallium-rich conditions. Therefore GaS will be prone to form VS donor defect. However, GaS has the highest formation energy, whereas Si has the lowest formation energy under sulfur-rich conditions, which proves that GaS will easily form Si acceptor defect.


image file: c5ra08329j-f6.tif
Fig. 6 Calculated formation energy of vacancy VGa in GaS as a function of EF under Ga-rich and S-poor conditions (ΔμGa = 0 and ΔμS = −1.114 eV). The zero point of EF corresponds to the VBM. The dotted line corresponds to the CBM in the calculation.
Table 2 Calculated defect formation energies ΔH at EF = 0 for intrinsic defects in GaS monolayer under Ga-rich (ΔμGa = 0 and ΔμS = −1.114 eV) and S-rich conditions (ΔμS = 0 and ΔμGa = −1.114 eV)
Defects q ΔH (eV)
Ga-rich S-rich
Gai 1+ −0.980 0.134
0 1.618 2.733
GaS 2+ −2.359 −0.131
1+ −0.373 1.855
0 1.912 4.140
VS 2+ −1.499 −0.385
1+ −0.246 0.868
0 0.948 2.063
VGa 0 3.811 2.697
1− 3.918 2.804
2− 5.234 4.120
Si 0 2.073 0.959
1− 2.467 1.353
SGa 0 5.034 2.805
1− 6.067 3.839
2− 6.860 4.631
3− 8.126 5.897


Fig. 7 shows the formation energies of the intrinsic defects in monolayer GaS as a function of EF under different conditions. Fig. 7a shows the formation energies of the intrinsic defects as a function of EF under gallium-rich and sulfur-poor conditions. It is clearly seen that the GaS defect with q = 2+ have the lowest and most negative formation energy near the VBM. Therefore GaS will be prone to form GaS donor defect. As the concentration of electrons increases, EF will move upward. When EF is located on the crossing point between GaS (q = 2+) and Gai (q = 1+), the GaS monolayer will tend to form Gai donor defect. Eventually, EF will stay on the crossing point between Gai (q = 1+) and Si (q = 1−) and close to the CBM. Therefore GaS will tend to form an n-type semiconductor under gallium-rich conditions.


image file: c5ra08329j-f7.tif
Fig. 7 Formation energies of intrinsic defects in GaS monolayer as a function of EF under (a) Ga-rich and S-poor conditions (ΔμGa = 0 and ΔμS = −1.114 eV) and (b) Ga-poor and S-rich conditions (ΔμS = 0 and ΔμGa = −1.114 eV). The zero point of EF corresponds to the VBM and the formation energies have been plotted up to the CBM with the experimental band gap (3.05 eV). The red horizontal line corresponds to a formation energy of zero. The slope of the line represents the charge state q labeled above each line.

Fig. 7b shows the formation energies of native defects as a function of EF under gallium-poor and sulfur-rich conditions. SGa defect with q = 3− have the lowest and most negative formation energies among all the intrinsic defects near the CBM. Consequently, GaS will tend to form SGa acceptor defects. As the concentration of cavity carriers increases, EF will remove downward. When EF is located on the crossing point between Si (q = 1−) and SGa (q = 3−), the GaS monolayer will be prone to form Si acceptor defects. Eventually, EF will remain on the crossing point between Gai (q = 1+) and Si (q = 1−) and close to the VBM. Therefore GaS tends to form a p-type semiconductor under sulfur-rich conditions.

Table 3 lists the transition levels ε(q/q′) of each point defect with respect to the VBM. To obtain n-type semiconductors, the primary donor defect should have a high transition level and low formation energy. Table 3 shows that the donor defect Gai has the highest transition level ε(1+/0) and is closest to the CBM. Thus Gai can provide abundant electron carriers to form n-type GaS monolayer. However, the formation energy of the donor defect GaS is lower than that of Gai in any case when EF is close to the VBM; therefore GaS monolayer does not tend to form Gai donor defect (Fig. 7a). Although GaS defect has a transition energy ε(1+/0) lower only than that of Gai, in view of the fact that GaS tends to form GaS defect more easily, GaS becomes the primary defect in n-type GaS despite its lower transition energy relative to Gai. We deduced that, to obtain an n-type semiconductor, the GaS monolayer should be prepared under gallium-rich conditions. On the other hand, to obtain p-type semiconductors, the dominant acceptor defect should have a low transition level and low formation energy. Table 3 shows that the acceptor defect VGa with the lowest transition energy level ε(0/1−) is closest to the VBM. This indicates that the VGa defect can provide p-type GaS with sufficient concentrations of cavities. However, the formation energy of the acceptor defect SGa is lowest among all the acceptor defects in any case when EF is close to the CBM; accordingly, the GaS monolayer does not tend to form VGa acceptor defect (Fig. 7b). Although the SGa defect has a transition energy ε(0/2−) higher than that of VGa, because GaS tends to form SGa defect more easily, SGa is the primary defect in p-type GaS despite its higher transition energy relative to VGa. From this discussion, we can predict that, to achieve a p-type GaS monolayer, preparation should take place under sulfur-rich conditions.

Table 3 Calculated transition energy levels ε(q/q′) of intrinsic defects in GaS monolayer with respect to the VBM
Defects q/q ε(q/q′) (eV)
Gai 1+/0 2.598
GaS 2+/1+ 1.986
1+/0 2.285
VS 2+/0 1.224
VGa 0/1− 0.107
1−/2− 1.316
Si 0/1− 0.394
SGa 0/2− 0.913
2−/3− 1.266


IV. Conclusions

We have studied the electronic and magnetic properties of the intrinsic point defects VGa, VS, GaS, SGa, Gai and Si in GaS monolayer and bulk sample on the basis of first-principles calculations. We also investigated the formation energies and transition energy levels of these defects in the monolayer. In the GaS monolayer, the impurity states exist in the band gaps of all the defects except the interstitial Si. Furthermore, half-metallic behavior can be attained in the presence of VGa and Gai. Although monolayers with VGa, GaS, SGa and Gai defects and bulk samples with VGa, GaS and SGa defects gain a total magnetic moment of 1.0 μB, monolayers with VS and Si and bulk sample with Gai defect are spin-unpolarized. Monolayer GaS will tend to form an n-type semiconductor under gallium-rich condition, but will tend to form a p-type semiconductor under sulfur-rich condition. Among the possible intrinsic defects, GaS and SGa are considered to be appropriate n- and p-type defects, respectively. These results provide an effective way to regulate the structural and electronic properties of GaS monolayer.

Author contributions

H.C. performed the density functional theory calculations. H.C., Y.L. and L.H. wrote the manuscript. J.L. guided the work. All authors have read the manuscript.

Acknowledgements

This work was supported by the National Science Foundation of China under Grant no. 91233120 and the National Basic Research Program of China (2011CB921901).

Notes and references

  1. A. K. Geim and K. S. Novoselov, Nat. Mater., 2007, 6, 183 CrossRef CAS PubMed.
  2. A. K. Geim, Science, 2009, 324, 1530 CrossRef CAS PubMed.
  3. C. N. R. Rao, A. K. Sood, K. S. Subrahmanyam and A. Govindaraj, Angew. Chem., Int. Ed., 2009, 48, 7752 CrossRef CAS PubMed.
  4. C. N. R. Rao, A. K. Sood, R. Voggu and K. S. Subrahmanyam, J. Phys. Chem. Lett., 2010, 1, 572 CrossRef CAS.
  5. A. Kuhn, A. Bourdon, J. Rigoult and A. Rimsky, Phys. Rev. B: Condens. Matter Mater. Phys., 1982, 25, 4081 CrossRef CAS.
  6. M. Gatulle, M. Fischer and A. Chevy, Phys. Status Solidi B, 1983, 119, 327 CrossRef CAS PubMed.
  7. M. Gatulle and M. Fischer, Phys. Status Solidi B, 1984, 121, 59 CrossRef CAS PubMed.
  8. R. S. Madatov, A. I. Najafov, T. B. Tagiev and A. R. Mobili, Inorg. Mater., 2008, 44, 100 CrossRef CAS.
  9. G. Shen, D. Chen, P.-C. Chen and C. Zhou, ACS Nano, 2009, 3, 1115 CrossRef CAS PubMed.
  10. G. Sinha, S. K. Panda, A. Datta, P. G. Chavan, D. R. Shinde, M. A. More, D. S. Joag and A. Patra, ACS Appl. Mater. Interfaces, 2011, 3, 2130 CAS.
  11. Q. S. Xin, S. Conrad and X. Y. Zhu, Appl. Phys. Lett., 1996, 69, 1244 CrossRef CAS PubMed.
  12. F. J. Manjón, A. Segura and V. Muñoz, J. Appl. Phys., 1997, 81, 6651 CrossRef PubMed.
  13. D. J. Late, B. Liu, J. Luo, A. Yan, H. S. S. R. Matte, M. Grayson, C. N. R. Rao and V. P. Dravid, Adv. Mater., 2012, 24, 3549 CrossRef CAS PubMed.
  14. S. Yang, Y. Li, X. Wang, N. Huo, J.-B. Xia, S.-S. Li and J. Li, Nanoscale, 2014, 6, 2582 RSC.
  15. D. J. Late, B. Liu, H. S. S. R. Matte, C. N. R. Rao and V. P. Dravid, Adv. Funct. Mater., 2012, 22, 1894 CrossRef CAS PubMed.
  16. V. Zólyomi, N. D. Drummond and V. I. Fal'ko, Phys. Rev. B: Condens. Matter Mater. Phys., 2013, 87, 195403 CrossRef.
  17. Y. Ma, Y. Dai, M. Guo, L. Yu and B. Huang, Phys. Chem. Chem. Phys., 2013, 15, 7098 RSC.
  18. G. S. Orudzhev and E. K. Kasumova, Phys. Solid State, 2014, 56, 619 CrossRef CAS.
  19. Y. Zhou, S. Li, W. Zhou, X. Zu and F. Gao, Sci. Rep., 2014, 4, 5773 CAS.
  20. Y. Ding and Y. Wang, Appl. Phys. Lett., 2013, 103, 043114 CrossRef PubMed.
  21. B. Wen, R. Melnik, S. Yao and T. Li, Mater. Sci. Semicond. Process., 2010, 13, 295 CrossRef CAS PubMed.
  22. Z. Zhu, Y. Cheng and U. Schwingenschlögl, Phys. Rev. Lett., 2012, 108, 266805 CrossRef.
  23. L. Huang, Z. Chen and J. Li, RSC Adv., 2015, 5, 5788 RSC.
  24. J. P. Perdew, K. Burke and M. Ernzerhof, Phys. Rev. Lett., 1996, 77, 3865 CrossRef CAS.
  25. P. E. Blöchl, Phys. Rev. B: Condens. Matter Mater. Phys., 1994, 50, 17953 CrossRef.
  26. G. Kresse and J. Hafner, Phys. Rev. B: Condens. Matter Mater. Phys., 1993, 47, 558 CrossRef CAS.
  27. G. Kresse and J. Furthmüller, Phys. Rev. B: Condens. Matter Mater. Phys., 1996, 54, 11169 CrossRef CAS.
  28. J. Heyd and G. E. Scuseria, J. Chem. Phys., 2003, 118, 8207 CrossRef CAS PubMed.
  29. J. Heyd, G. E. Scuseria and M. Ernzerhof, J. Chem. Phys., 2006, 124, 219906 CrossRef PubMed.
  30. H. J. Monkhorst and J. D. Pack, Phys. Rev. B: Condens. Matter Mater. Phys., 1976, 13, 5188 CrossRef.
  31. S. B. Zhang and S.-H. Wei, Appl. Phys. Lett., 2002, 80, 1376 CrossRef CAS PubMed.
  32. S. B. Zhang and J. E. Northhrup, Phys. Rev. Lett., 1991, 67, 2339 CrossRef CAS.
  33. G. A. Baraff and M. Schlüter, Phys. Rev. Lett., 1985, 55, 1327 CrossRef CAS.
  34. S. B. Zhang and J. E. Northrup, Phys. Rev. Lett., 1991, 67, 2339 CrossRef CAS.
  35. D. B. Laks, C. G. Van de Walle, G. F. Neumark, P. E. Blöchl and S. T. Pantelides, Phys. Rev. B: Condens. Matter Mater. Phys., 1992, 45, 10965 CrossRef CAS.
  36. C. Persson, Y.-J. Zhao, S. Lany and A. Zunger, Phys. Rev. B: Condens. Matter Mater. Phys., 2005, 72, 035211 CrossRef.
  37. H. Chen, C.-Y. Wang, J.-T. Wang, X.-P. Hu and S.-X. Zhou, J. Appl. Phys., 2012, 112, 084513 CrossRef PubMed.
  38. J. He, K. Wu, R. Sa, Q. Li and Y. Wei, Appl. Phys. Lett., 2010, 96, 082504 CrossRef PubMed.
  39. S.-H. Wei, Comput. Mater. Sci., 2004, 30, 337 CrossRef CAS PubMed.
  40. J. Pohl and K. Albe, J. Appl. Phys., 2010, 108, 023509 CrossRef PubMed.
  41. S.-H. Wei and S. B. Zhang, Phys. Rev. B: Condens. Matter Mater. Phys., 2002, 66, 155211 CrossRef.
  42. A. Janotti and C. G. Van de Walle, Phys. Rev. B: Condens. Matter Mater. Phys., 2007, 76, 165202 CrossRef.
  43. A. Alkauskas, P. Broqvist and A. Pasquarello, Phys. Status Solidi B, 2011, 248, 775 CrossRef CAS PubMed.
  44. H. Chen, C.-Y. Wang, J.-T. Wang, Y. Wu and S.-X. Zhou, Phys. B, 2013, 413, 116 CrossRef CAS PubMed.
  45. Y. Yan and S.-H. Wei, Phys. Status Solidi B, 2008, 245, 641 CrossRef CAS PubMed.
  46. P. Hu, L. Wang, M. Yoon, J. Zhang, W. Feng, X. Wang, Z. Wen, J. C. Idrobo, Y. Miyamoto, D. B. Geohegan and K. Xiao, Nano Lett., 2013, 13, 1649 CAS.
  47. C. H. Ho and S. L. Lin, J. Appl. Phys., 2006, 100, 083508 CrossRef PubMed.
  48. G. Henkelman, A. Arnaldsson and H. Jónsson, Comput. Mater. Sci., 2006, 36, 354 CrossRef PubMed.
  49. Q. Yue, S. Chang, S. Qin and J. Li, Phys. Lett. A, 2013, 377, 1362 CrossRef CAS PubMed.

This journal is © The Royal Society of Chemistry 2015
Click here to see how this site uses Cookies. View our privacy policy here.