First-principles study of half-fluorinated silicene sheets

Xiao Wang*a, Huangzhong Liub and Shan-Tung Tuc
aSchool of Science, East China University of Science and Technology, Shanghai 200237, China. E-mail: laricswang@gmail.com
bDepartment of Equipment Economics Management, PLA Military Economics Academy, Wuhan 430035, China
cKey Laboratory of Pressure Systems and Safety, Ministry of Education, East China University of Science and Technology, Shanghai 200237, China

Received 13th October 2014 , Accepted 15th December 2014

First published on 15th December 2014


Abstract

The structural, electronic and magnetic properties of half-fluorinated silicene sheets are investigated using first-principles simulation. Three conformers of half-fluorinated silicene are studied and their properties are compared. Half-fluorinated silicene sheets with zigzag, boat-like or chair-like configurations are confirmed as dynamically stable based on phonon calculations. Upon the adsorption of fluorine, energy gaps open in both zigzag and boat-like conformations. They are found to be direct-gap semiconductors. The half-fluorinated silicene with chair-like configuration shows an antiferromagnetic behavior which is mainly induced by the un-fluorinated Si atoms. Furthermore, when isotropic strain is uniformly exerted onto chair-like half-fluorinated silicene, the energy difference between ferromagnetism states and antiferromagnetism states decreases with increasing compression strain from 0% to −6% or increasing tension strain from 0% to 6%. These results demonstrate that fluorination with different atomic configurations is an efficient way to tune the electronic structures and properties of silicene sheets and highlight the potential of half-fluorinated silicene for spintronics.


1. Introduction

Graphene, a two-dimensional (2D) honeycomb network of carbon atoms, has attracted many interests since Novoselov et al. first fabricated it in 2004.1–3 It has zero-band-gap semi-metallic behavior with a Dirac-like electronic excitation and shows many remarkable properties such as high carrier mobility, quantum Hall effect for relatively low magnetic fields at room temperature and so on.2,4 Because of these outstanding properties, graphene has been suggested to be a very promising candidate of the future electronic materials. To expand the applications of graphene-based materials into nanoelectronics and spintronics, fluorination is a promising and effective way to open the band gap and induce magnetism of it. Fluorine (F) atoms are introduced into graphene in the form of C–F covalent bond by XeF2 fluorination.5 Experimental and theoretical studies have indicated that the band gap of fluorinated graphene can be varied from 0 eV to about 3 eV with changing degrees of fluorination.6–10 More interestingly, magnetism can be induced in half-fluorinated graphene which is an antiferromagnetic (AFM) semiconductor.11

The impressive progress in graphene has motivated the exploration of other 2D atomic based materials, such as BN and GaN sheet.12,13 Researches on fluorinated BN and GaN sheets indicated that half-fluorinated BN sheet becomes a direct band gap semiconductor and is an antiferromagnet,14 while half-fluorinated GaN sheet shows ferromagnetism (FM) behavior.11 As the counterpart of graphene, the 2D hexagonal lattices of silicon (Si), so-called silicene, recently were synthesized in the experiments.15–21 Various studies have revealed that silicene not only has a similar electronic structure with graphene, but also has some unique features such as a large spin–orbit gap at the Dirac Point,22 experimentally accessible quantum spin Hall effect,23 electrically tunable band gap24 and the emergence of a valley-polarized metal phase.25 Because silicene sheets have recently been realized, it is highly desirable to investigate if fluorination of silicene can induce some new effects. Ding et al.26 have reported that after full fluorination, a direct band gap can be opened in the silicene fluoride and the gap values are continuously modulated by the strain. However, little work has been done to investigate the properties of partially fluorinated silicene, for example, the silicene with half of Si atoms saturated with F atoms. In addition, based on the previous studies, both silicene and full fluorinated silicene sheets are known to be nonmagnetic. Is it possible that half-fluorinated silicene can also be magnetic like half-fluorinated graphene? And does the arrangement of F atoms influence the properties of half-fluorinated silicene?

To explore these problems, in this work, we performed detailed first-principles calculations to investigate the structural, electronic and magnetic properties of the half-fluorinated silicene with three different conformers. On the basis of density functional theory (DFT) computations, it is found that the arrangement of the doped F atoms has significant influence on the electronic structures of silicene. Only half-fluorinated silicene with chair-like configuration shows magnetic behavior while zigzag and boat-like conformations become direct band-gap semiconductors with nonmagnetic. In addition, the strain effect on the magnetic properties of half-fluorinated silicene is also studied here.

2. Computational methods

The present calculations have been performed by using the Castep package,27 which is based on density functional theory (DFT). The exchange–correlation functional is treated within the generalized gradient approximation (GGA), in the form of Perdew–Burke–Ernzerhof (PBE) functional.28 Ultrasoft pseudopotential29 is adopted for the spin-polarized computation and the plane-wave cutoff energy is set to be 450 eV. In the pseudopotentials, the 3s, 3p shells of Si atom and the 2s, 2p shells of F atom are treated as valence shells. Supercells are used to simulate the isolated nanostructures, and the distance between images is 20 Å to eliminate the interactions between the adjacent Si layers. The Brillouin zone is represented by the set of 10 × 10 × 1 k-points30 for the geometry optimizations, and 15 × 15 × 1 k-points are used to obtain the density of states (DOS). All the lattice constants and atom coordinates are optimized until the convergence of the force on each atom is less than 0.01 eV Å−1. Also, the hybrid HSE03 functional31 calculation by using norm-conserving pseudopotentials32 is performed to obtain more correct description of band gaps due to the well-known problem of GGA calculations on evaluating the band gaps.

To determine the stability of the half-fluorinated silicene, the formation energy of the fluorinated single layer is estimated by calculating Ef = (EtotalEpureNFEF)/NF, where Etotal is the energy of the half-fluorinated silicene, Epure is that of the pristine silicene, EF is the binding energy per atom of an F2 molecule, and NF is the number of adsorbed F atoms.

3. Results and discussion

3.1 Pristine silicene

The optimized geometric structure of pristine silicene is shown in Fig. 1(a). The lattice parameter of relaxed silicene equals 3.866 Å and the Si–Si bond length is calculated to be d = 2.280 Å, consistent with previous results.33,34 Compared with graphene, which has very strong π bonding and planar geometry, the larger Si–Si interatomic distance in silicene weakens the π–π overlaps and dehybridizes the sp2 states which results in a low-buckled structure with h = 0.45 Å (Fig. 1(b)). Silicene is also a semimetal (gapless semiconductor) like graphene35,36 shown in Fig. 2(a), in which a Dirac cone exists and is formed by the crossing of the bonding π and antibonding π* bands at k points in the hexagonal Brillouin zone. In addition, the projected DOS (PDOS) in Fig. 2(b) points out again that in silicene, the Dirac cone exists at the Fermi level and results from the 3pz orbitals of Si atoms.
image file: c4ra12257g-f1.tif
Fig. 1 Optimized geometric structures of silicene from (a) the top view and (b) the side view. Silicon atoms are shown in yellow.

image file: c4ra12257g-f2.tif
Fig. 2 (a) Band structure of pristine silicene calculated after structural relaxation along high symmetry directions and (b) its PDOS. The Fermi level is set to 0 eV and displayed in dots.

3.2 Half-fluorinated silicene sheets

When silicene is exposed to gaseous fluorine, three possible atomic configurations of half-fluorinated silicene are considered: zigzag conformer with F atoms absorbed on the zigzag chains alternately (labeled as zigzag-F@Si2 and shown in Fig. 3(a)), boat-like conformer with F atoms distributed in pairs (labeled as boat-F@Si2 and shown in Fig. 3(b)), and chair-like conformer with F atoms doped alternately (labeled as chair-F@Si2 and shown in Fig. 3(c)). In these configurations, F atoms are only doped on one side of the silicene sheet. We defined Si1 atoms as Si atoms remain unsaturated and Si2 atoms as Si atoms doped by F atoms here. The optimized structural parameters and the formation energies of three conformers are listed in Table 1. Compared with pristine silicene, the bond lengths of Si–Si bonds in fluorinated silicene have a varying degree of increase due to the formation of Si–F bonds. And the buckling heights hbuckle of Si atoms increase significantly, especially for zigzag-F@Si2 sheet. The formation energies of these half-fluorinated silicene are all negative, indicating that the fluorinations of silicene are exothermic reactions and the corresponding half-fluorinated silicene structures are stable. To verify the stability of half-fluorinated silicene, phonon calculations are also carried out and no imaginary frequencies are found in all the three conformers, indicating that the structures are dynamically stable. The phonon dispersion curves are plotted in Fig. 4. The top frequencies are related with the Si–F bonds, which could be useful in characterizing these conformations. Among the three conformers, the zigzag configuration is the most stable one and the energies follow the order of Ef(zigzag) < Ef(boat-like) < Ef(chair-like) as shown in Table 1.
image file: c4ra12257g-f3.tif
Fig. 3 Top and side views of optimized structures of (a) zigzag, (b) boat-like, and (c) chair-like half-fluorinated silicene, respectively. Fluorine atoms are shown in blue.
Table 1 Lattice constants, bond lengths (in Å) and the formation energies (in eV per atom) of half-fluorinated silicene sheets. γ is the angle between the two lattice vectors
  Zigzag Boat-like Chair-like
a 3.924 3.990 3.896
b 6.796 6.911
γ 90 90 60
hbuckle 1.216 0.801 0.753
dSi1–Si1 2.296 2.231
dSi1–Si2 2.383 2.397 2.350
dSi2–Si2 2.391 2.448
dSi2–F 1.636 1.634 1.634
Ef −3.575 −3.519 −3.311



image file: c4ra12257g-f4.tif
Fig. 4 The phonon dispersion curves for three conformations of half-fluorinated silicene. (a) Zigzag; (b) boat-like and (c) chair-like.

In general, there are three kinds of Si–Si bonds existed in three conformations of half-fluorinated silicene: Si1–Si1, Si2–Si2 and Si1–Si2. To understand the charge transfer and bonding characters, we take the most stable conformer, zigzag half-fluorinated silicene sheet, as an example and present its electron density difference in Fig. 5. The electron density difference Δρ is defined as the difference between the total charge density of half-fluorinated silicene system and the superposition of independent charge density of silicene and F atoms.

Δρ = ρtotal − (ρsilicene + ∑ρFi)
where ρtotal, ρsilicene and ρF denote electron density of zigzag half-fluorinated silicene sheet, pristine silicene and each F atoms, respectively. Loss of electrons is indicated in red, while electron enrichment is indicated in blue. When F atom doped on silicene sheet, a charge transfer from Si atom to F atom is clearly shown in Fig. 5 which leaves Si atoms positively charged. The Si–Si bonds are strongly covalent while the Si–F bonds are primarily ionic mixed with partial covalent. Furthermore, it is worth noting that in Table 1 the Si1–Si1 bonds in zigzag-F@Si2 and boat-F@Si2 sheets are all shorter than the corresponding Si1–Si2 bonds and Si2–Si2 bonds. This is because that the pz orbitals on the Si2 atoms are saturated by F atoms, which destroys the formation of π bonding network in the Si1–Si2 bond and Si2–Si2 bonds, while besides the strong σ bonds between the Si1–Si1 atoms, the π bonds between the unpaired pz electrons of the Si1 atoms are also formed, thus makes the Si1–Si1 bond even shorter.


image file: c4ra12257g-f5.tif
Fig. 5 Electron density difference for three kinds of Si–Si bonds in zigzag-F@Si2 sheet: (a) Si1–Si1 bond; (b) Si2–Si2 bond; (c) Si1–Si2 bond. Loss of electrons is indicated in red, while electron enrichment is indicated in blue.

The sp3 hybridization of fluorine also affects the electronic structures of silicene. For the half-fluorinated silicene with zigzag and boat-like conformations, it is represented in Table 2 and Fig. 6 that they become direct band-gap semiconductors, both the valence band maximum (VBM) and conduction band minimum (CBM) are located at the Γ point, and small gaps of 0.193 eV and 0.396 eV are opened, respectively. When Si atoms are exposed to F atoms, the 3pz orbitals of Si atoms bond with 2p orbitals of F atoms, thus the weak coupling between Si atoms are broken which leads to the semiconducting nature of the systems. In Fig. 7, the CBM of boat-F@Si2 sheet is composed of the pz orbitals of unsaturated Si (Si1) atoms and anti-bonding states of Si–F bonds, while the VBM is composed of the σ bonding states of Si1–Si2 bonds along the zigzag lines. When the half-fluorinated silicene has zigzag configuration, its VBM in Fig. 8 is not only composed of the σ bonding states of Si1–Si2 bonds, but also the week π bonding between the unsaturated Si1 atoms along the zigzag chain which makes a direct gap feature kept but with a bit smaller gap of 0.193 eV compared with that of boat-F@Si2 sheet. Additionally, to circumvent the problem of GGA calculations which usually underestimates the energy gap of semiconductors, the hybrid HSE03 calculations are performed to compare with the GGA results. The calculated energy gaps are also given in Table 2. For the systems studied here, the GGA energy gaps are found to be typically 40–55% smaller than the HSE results. However, there is no difference between the GGA and HSE evaluations on the type of the band gaps.

Table 2 Energy gaps of half-fluorinated silicene, computed by using GGA and HSE03 hybrid functionals
Conformers Energy gap by GGA (eV) Energy gap by HSE (eV)
Zigzag-F@Si2 0.193 (direct) 0.411 (direct)
Boat-F@Si2 0.396 (direct) 0.824 (direct)
Chair-F@Si2 0.243 (direct) 0.468 (direct)



image file: c4ra12257g-f6.tif
Fig. 6 Band structures of (a) zigzag and (b) boat-like half-fluorinated silicene sheets, calculated using PBE functional. The Fermi levels are set as 0 eV and shown in dots.

image file: c4ra12257g-f7.tif
Fig. 7 The PDOS (a) and (b) partial charge densities of VBM and CBM for the boat-like conformation of half-fluorinated silicene. The Fermi levels are set as 0 eV and shown in dots.

image file: c4ra12257g-f8.tif
Fig. 8 The PDOS (a) and (b) partial charge densities of VBM and CBM for the zigzag conformation of half-fluorinated silicene. The Fermi levels are set as 0 eV and shown in dots.

Although the boat-like and zigzag conformations of half-fluorinated silicene are nonmagnetic, Mulliken analysis shows that in the case of chair-F@Si2 sheet, it is spin-polarized with a local magnetic moment of about 1 μB per unit cell, similar with half-hydrogenated silicene or half-brominated silicene.37,38 In order to study the preferred coupling, different magnetic configurations are presented in Fig. 9: (1) nonmagnetic (NM); (2) FM coupling; (3) AFM coupling states within four types (A-AFM, G-AFM, H-AFM and P-AFM). Both PBE and HSE results are summarized in Table 3 and show that the A-AFM coupling is the most stable one and lies 0.104 and 0.406 eV (0.108 and 0.424 eV calculated by HSE) lower per unit cell in energy than that of FM and NM states, respectively, indicating that chair-F@Si2 exhibits an AFM behavior. Also, the energies of G, H and P-type AFM states calculated by PBE functional are 0.009, 0.054 and 0.053 eV higher per unit cell than that of A-AFM coupling, respectively. Thus, the discussions in the following are focused on chair-F@Si2 with A-AFM character and directly label it as AFM. The corresponding band structure of chair-F@Si2 in AFM state is plotted in Fig. 10, showing that the system is a direct semiconductor with band gap of 0.243 eV by GGA calculation. Detailed analysis of PDOS (Fig. 11) reveals that the magnetism is mainly contributed by the unpaired 3p orbitals of unsaturated Si1 atoms. In chair-F@Si2 sheet, Si2 atoms are bonded with F atoms forming strong σ-bonds; and the π-bonding network in the silicene sheet is completely broken, leaving 3pz electrons of Si1 atoms unpaired and localized. It is clearly shown in Fig. 11 that spin splits near the Fermi level and thus induces magnetism. In fact, Mulliken population analysis shows that per Si1 atom carried a magnetic moment of about 0.84 μB, while very small magnetic moment was found on F and Si2 atoms (about 0.02 and 0.01 μB per atom, respectively). The corresponding spatial spin-density distributions are illustrated in Fig. 12.


image file: c4ra12257g-f9.tif
Fig. 9 Different types of AFM coupling states of chair-F@Si2 sheet.
Table 3 The relative energies per unit cell (in eV) and calculated magnetic moment (in μB) for the atoms of chair-F@Si2 in different magnetic configurations
Coupling ΔEPBE ΔEHSE MSi1 MSi2 MF
NM 0.406 0.424 0 0 0
FM 0.104 0.108 0.94 0.02 0.04
A-AFM 0 0 ±0.84 ±0.01 ±0.02
G-AFM 0.009 0.002 ±0.84 ±0.01 ±0.02
H-AFM 0.054 0.050 ±0.90 ±0.03 ±0.06
P-AFM 0.053 0.053 ±0.90 ±0.02 ±0.03



image file: c4ra12257g-f10.tif
Fig. 10 Band structure of chair-F@Si2 sheet in A-AFM state, calculated using GGA functional. The arrows represent the spin direction.

image file: c4ra12257g-f11.tif
Fig. 11 The PDOS of chair-F@Si2 sheet in A-AFM state. The vertical dotted line represents the Fermi level.

image file: c4ra12257g-f12.tif
Fig. 12 Spin density of chair-F@Si2 sheet in A-AFM state. Purple and gray indicate the positive and negative values of spin, respectively.

3.3 Applied strain on chair-like half-fluorinated silicene sheet

Previous studies have indicated that the spin polarization exhibits remarkable radius dependence in F-BNNTs,39,40 and increasing local tube curvature can enhance the relatively weak FM order in F-BNNTs. Another theoretical study on half-fluorinated BN sheet and graphene11 showed that the energy difference between ferromagnetic and antiferromagnetic couplings decreases significantly with strain increasing. Motivated by these results, a nanomechanical modulation of strain may sensitively influence the spin order of half-fluorinated silicene. Thus, we investigate the strain dependence of spin order in chair-F@Si2 sheet by varying the isotropic strain from −6% to 6%, in which all crystal symmetries and honeycomb-like structures are maintained. The in-plane tensile or compression strain is uniformly applied along lattice directions. The isotropic strain is defined as ε = Δa/a0, where the lattice constants of the unstrained and strained supercell equal to a0 and a = Δa + a0, respectively. The stretching or compressing of the chair-F@Si2 sheet is achieved by first elongating or shortening the optimized lattice constant a0 to a and re-optimized. Then the corresponding energy is calculated with the elongated or shortened lattice constant fixed. Fig. 13 shows the variation in energy difference per unit cell between FM and AFM (A type) order ΔE of chair-F@Si2 sheet with strain. Different with half-fluorinated graphene, it is found that for half-fluorinated silicene, ΔE dramatically decreases with increasing isotropic compression from 0% to −6% and also decreases gently with increasing isotropic tension from 0% to 6%. It indicates that the interaction between the moments keeps AFM coupling independent of the tension/compression, which reveals that the AFM spin order in chair-like half-fluorinated silicene is robust within a wide range of strain.
image file: c4ra12257g-f13.tif
Fig. 13 Strain dependence of the energy difference per unit cell between FM and AFM order of chair-F@Si2 sheet.

According to the previous study by Ma et al.,11 the magnetism of half-decorated 2D-sheets can be determined by the competition of two distinct interactions, through-bond interaction and p–p direct interaction. The through-bond interaction is defined as a kind of interactions that a atom with spin up (spin down) density induces a spin down (spin up) density on the adjacent b atom which directly bonds to it. It is an indirect interaction mediated by b atom which may lead to FM spin coupling even in a long range. In the situation of half-decorated silicene we studied here, a atom refers to unsaturated Si1 atom and b atom means Si2 atom doped by F atom. In contrast, the p–p direct spin polarization can induce the spin of the dangling bonds of a atoms into AFM coupling. In this kind of interaction, an a atom with spin up (spin down) density induces a spin down (spin up) density on the nearest-neighboring a atom directly, without mediated by b atom. In chair-F@Si2 sheet, the strengthened bulking structure leads to the much larger spatial extension of 3p states of Si1 atoms, which promotes direct p–p interaction and results in a more stable AFM coupling than the FM coupling. When strain is exerted on the sheet, the distance between two neighboring Si1 atoms and the bond length of adjacent Si1–Si2 increase with increasing tensile strain (or decreasing compression strain), which results in the reduction of both through-bond and p–p direct interactions. The elongation of Si1–Si2 bond reduces bond energy and induces a little more spin-polarized electronic states (Fig. 14). The magnetic moment MSi1 of the strained chair-F@Si2 sheet with ε = 6% slightly increases to 0.9 μB. At the meantime, the decrease of the p–p interaction is a little larger than that of the through-bond interaction. As a result, for the chair-F@Si2 sheet under tensile strain, the interaction between the moments keeps AFM coupling. On the other hand, PDOS of chair-F@Si2 sheet under 6% compression demonstrates that the strengthened interaction between the adjacent Si atoms weakens the spin split of Si1 3p states, thus some unoccupied states sink below the Fermi level and become the occupied states. The magnetic moment MSi1 is quenched from 0.84 μB to 0.38 μB, indicating that the magnetic coupling under such conditions is quenched. Although the AFM ordering is weakened for compression stress on chair-F@Si2 sheet, however, the p–p interaction still plays a dominant role in spin polarization. As compared to the distance between two neighboring Si1 atoms, the bond length of Si1–Si2 decreases more obviously when the sheet is compressed. Thus, the increase of the through-bond interaction is larger than that of the p–p direct interaction with increasing compression strain, resulting in the relative decrease of the p–p direct interaction. This is why the energy difference ΔE decreases with increasing isotropic compression from 0% to −6%, also decreases with increasing isotropic strain from 0% to 6%. The variation of spin polarization in half-fluorinated silicene under strain may find applications in nanodevices, such as a mechanical switch for spin-polarized transport or many bistable devices.


image file: c4ra12257g-f14.tif
Fig. 14 PDOS of chair-F@Si2 under −6%, 0% and 6% strain, respectively. The vertical dotted line represents the Fermi level.

4. Conclusions

In conclusion, using first-principles calculations, we systematically investigate the structures and properties of half-fluorinated silicene. The atomic configurations have great effect on their electronic structures and magnetism. The zigzag conformer is also the most stable configuration of half-fluorinated silicene, same to the situation of full fluorination of silicene. And direct band gaps are opened in the zigzag-F@Si2 and boat-F@Si2 sheets with values of 0.411 and 0.824 eV obtained from the HSE03 functional, respectively. Only the half-fluorinated silicene with chair-like configuration shows antiferromagnetism. In chair-F@Si2 sheet, electrons in Si1 atoms are unpaired and localized, this results in spin polarization. Furthermore, the in-plane strain can be used to tune the relative stability of FM and AFM states of chair-F@Si2 sheet due to the combined effects of both through-bond and p–p direct interaction. Once combined with advanced Si nanotechnology, these predicted properties may make silicene be useful as a promising nanoscale technological application in electron component and spintronics.

Acknowledgements

The authors gratefully acknowledge financial support from NSFC (no. 21303054) and China Postdoctoral Science Foundation (2013M540332). All the computation simulation was undertaken with the resources provided from the High Performance Computing Center of East China University of Science and Technology.

References

  1. A. K. Geim and K. S. Novoselov, Nat. Mater., 2007, 6, 183 CrossRef CAS PubMed.
  2. A. H. Castro Neto, F. Guinea, N. M. R. Peres, K. S. Novoselov and A. K. Geim, Rev. Mod. Phys., 2009, 81, 109 CrossRef CAS.
  3. K. S. Novoselov, A. K. Geim, S. V. Morozov, D. Jiang, Y. Zhang, S. V. Dubonos, I. V. Grigorieva and A. A. Firsov, Science, 2004, 306, 666 CrossRef CAS PubMed.
  4. D. S. L. Abergel, V. Apalkov, J. Berashevich, K. Ziegler and T. Chakraborty, Adv. Phys., 2010, 59, 261 CrossRef CAS.
  5. R. R. Nair, W. Ren, R. Jalil, I. Riaz, V. G. Kravets, L. Britnell, P. Blake, F. Schedin, A. S. Mayorov, S. J. Yuan, M. I. Katsnelson, H. M. Cheng, W. Strupinski, L. G. Bulusheva, A. V. Okotrub, I. V. Grigorieva, A. N. Grigorenko, K. S. Novoselov and A. K. Geim, Small, 2010, 6, 2877 CrossRef CAS PubMed.
  6. J. T. Robinson, J. S. Burgess, C. E. Junkermeier, S. C. Badescu, T. L. Reinecke, F. K. Perkins, M. K. Zalalutdniov, J. W. Baldwin, J. C. Culbertson, P. E. Sheehan and E. S. Snow, Nano Lett., 2010, 10, 3001 CrossRef CAS PubMed.
  7. D. K. Samarakoon, Z. Chen, C. Nicolas and X. Q. Wang, Small, 2011, 7, 965 CrossRef CAS PubMed.
  8. S. H. Cheng, K. Zou, F. Okino, H. R. Gutierrez, A. Gupta, N. Shen, P. C. Eklund, J. O. Sofo and J. Zhu, Phys. Rev. B: Condens. Matter Mater. Phys., 2010, 81, 205435 CrossRef.
  9. K. J. Jeon, Z. Lee, E. Pollak, L. Moreschini, A. Bostwick, C. M. Park, R. Mendelsberg, V. Radmilovic, R. Kostecki, T. J. Richardson and E. Rotenberg, ACS Nano, 2011, 5, 1042 CrossRef CAS PubMed.
  10. J. C. Garcia, D. D. B. Lima, L. V. C. Assali and J. F. Justo, J. Phys. Chem. C, 2011, 115, 13242 CAS.
  11. Y. D. Ma, Y. Dai, M. Guo, C. W. Niu, L. Yu and B. B. Huang, Nanoscale, 2011, 3, 2301 RSC.
  12. J. C. Meyer, A. Chuvilin, G. Algara-Siller, J. Biskupek and U. Kaiser, Nano Lett., 2009, 9, 2683 CrossRef CAS PubMed.
  13. H. Şahin, S. Cahangirov, M. Topsakal, E. Bekaroglu, E. Akturk, R. T. Senger and S. Ciraci, Phys. Rev. B: Condens. Matter Mater. Phys., 2009, 80, 155453 CrossRef.
  14. J. Zhou, Q. Wang, Q. Sun and P. Jena, Phys. Rev. B: Condens. Matter Mater. Phys., 2010, 81, 085442 CrossRef.
  15. B. Lalmi, H. Oughaddou, H. Enriquez, A. Kara, S. Vizzini, B. Ealet and B. Aufray, Appl. Phys. Lett., 2010, 97, 223109 CrossRef PubMed.
  16. B. J. Feng, Z. J. Ding, S. Meng, Y. G. Yao, X. Y. He, P. Cheng, L. Chen and K. H. Wu, Nano Lett., 2012, 12, 3507 CrossRef CAS PubMed.
  17. P. Vogt, D. P. Padova, C. Quaresima, J. Avila, E. Frantzeskakis, M. C. Asensio, A. Resta, B. Ealet and G. Le Lay, Phys. Rev. Lett., 2012, 108, 155501 CrossRef.
  18. C. L. Lin, R. Arafune, K. Kawahara, N. Tsukahara, E. Minamitani, Y. Kim, N. Takagi and M. Kawai, Appl. Phys. Express, 2012, 5, 045802 CrossRef.
  19. H. Jamgotchian, Y. Colignon, N. Hamzaoui, B. Ealet, J. Y. Hoarau, B. Aufray and J. P. Biberian, J. Phys.: Condens. Matter, 2012, 24, 172001 CrossRef CAS PubMed.
  20. D. Chiappe, C. Grazianetti, G. Tallarida, M. Fanciulli and A. Molle, Adv. Mater., 2012, 24, 5088 CrossRef CAS PubMed.
  21. L. Meng, Y. L. Wang, L. Z. Zhang, S. X. Du, R. T. Wu, L. F. Li, Y. Zhang, G. Li, H. T. Zhou, W. A. Hofer and H. J. Gao, Nano Lett., 2013, 13, 685 CrossRef CAS PubMed.
  22. C. C. Liu, H. Jiang and Y. Yao, Phys. Rev. B: Condens. Matter Mater. Phys., 2011, 84, 195430 CrossRef.
  23. C. C. Liu, W. X. Feng and Y. G. Yao, Phys. Rev. Lett., 2011, 107, 076802 CrossRef.
  24. N. D. Drummond, V. Zólyomi and V. I. Fal'ko, Phys. Rev. B: Condens. Matter Mater. Phys., 2012, 87, 075423 CrossRef.
  25. M. Ezawa, Phys. Rev. Lett., 2012, 109, 055502 CrossRef.
  26. Y. Ding and Y. L. Wang, Appl. Phys. Lett., 2012, 100, 083102 CrossRef PubMed.
  27. M. D. Segall, P. J. D. Lindan, M. J. Probert, C. J. Pickard, P. J. Hasnip, S. J. Clark and M. C. Payne, J. Phys.: Condens. Matter, 2002, 14, 2717 CrossRef CAS.
  28. J. P. Perdew, K. Burke and M. Ernzerhof, Phys. Rev. Lett., 1996, 77, 3865 CrossRef CAS.
  29. D. Vanderbilt, Phys. Rev. B: Condens. Matter Mater. Phys., 1990, 41, 7892 CrossRef.
  30. H. J. Monkhorst and J. D. Pack, Phys. Rev. B: Solid State, 1976, 13, 5188 CrossRef.
  31. J. Heyd, G. E. Scuseria and M. Ernzerhof, J. Chem. Phys., 2003, 118, 8207 CrossRef CAS PubMed.
  32. D. R. Hamann, M. Schluter and C. Chiang, Phys. Rev. Lett., 1979, 43, 1494 CrossRef CAS.
  33. S. Lebègue and O. Eriksson, Phys. Rev. B: Condens. Matter Mater. Phys., 2009, 79, 115409 CrossRef.
  34. X. D. Li, J. T. Mullen, Z. H. Jin, K. M. Borysenko, M. B. Nardelli and W. K. Kim, Phys. Rev. B: Condens. Matter Mater. Phys., 2013, 87, 115418 CrossRef.
  35. Y. C. Cheng, Z. Y. Zhu and U. Schwingenschlögl, Europhys. Lett., 2011, 95, 17005 CrossRef.
  36. S. Cahangirov, M. Topsakal, E. Aktürk, H. Sahin and S. Ciraci, Phys. Rev. Lett., 2009, 102, 236804 CrossRef CAS.
  37. C. W. Zhang and S. S. Yan, J. Phys. Chem. C, 2012, 116, 4163 CAS.
  38. F. B. Zheng and C. W. Zhang, Nanoscale Res. Lett., 2012, 7, 422 CrossRef PubMed.
  39. Z. H. Zhang and W. L. Guo, J. Am. Chem. Soc., 2009, 131, 6874 CrossRef CAS PubMed.
  40. F. Li, Z. Zhu, X. Yao and G. Lu, Appl. Phys. Lett., 2008, 92, 102515 CrossRef PubMed.

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