Hydrogenation-induced large-gap quantum-spin-Hall insulator states in a germanium–tin dumbbell structure

Xin Chen, Linyang Li and Mingwen Zhao*
School of Physics, State Key Laboratory of Crystal Materials, Shandong University, Jinan, Shandong 250100, China. E-mail: zmw@sdu.edu.cn

Received 17th June 2015 , Accepted 20th August 2015

First published on 20th August 2015


Abstract

The quantum spin Hall (QSH) effect in two-dimensional topological insulators (2D TIs) is promising for building nanoscaled devices with low energy consumption. The 2D TIs with a bulk band gap larger than the atomic thermal motion energy at room temperature (∼26 meV) are essential for achieving room-temperature QSH effects. We proved from first-principles that this goal may be reached in a hydrogenated germanium–tin (Sn6Ge4H4) dumbbell (DB) structure, where spin–orbit coupling (SOC) opens a bulk band gap of 235 meV. The topological nontriviality is related to the band inversion of s–pxy of Sn atoms induced by surface hydrogenation and can be characterized by a topological invariant of Z2 = 1. This work offers a promising candidate material for achieving long-desired room-temperature QSH effects.


Introduction

Topological insulators (TIs) are a new quantum state of matter with a topologically nontrivial bulk band gap due to spin–orbit coupling (SOC) and gapless surface or edge states. Protected by time-reversal symmetry, low-energy scattering of the edge states in two-dimensional (2D) TIs (also called quantum spin Hall (QSH) insulators) is enjoined, resulting in dissipationless transport edge channels,1–9 which are quite promising for application in spintronics and quantum computations. To achieve such application, the bulk band gap in TIs should be larger than the atomic thermal motion energy. Although many materials have been predicted to be QSH insulators, only HgTe/CdTe10 and InAs/GaSb11 quantum wells were verified by transport experiments at extremely low temperature, due to their small bulk gap. Increasing the critical temperature is highly desirable for achieving the QSH effect at room-temperature. This has motivated an intensive search for large-gap 2D QSH insulators, such as metal–organic frameworks12–15 and those containing heavy metal atoms.16–20 Additionally, our recent work showed that tensile strain can serve as an effective means to drive a normal insulator to a QSH insulator.21

Inspired by the successful discovery and application of graphene, extensive research into the 2D group-IV materials has been conducted, and many exotic electronic properties have been revealed.22–29 For example, low-buckled (LB) silicene, germanene and stanene30 and their derivatives31,32 have been reported as QSH insulators with relatively strong SOC. Some of them have a bulk gap larger than the thermal motion energy at room temperature (∼26 meV). Apart from the LB configurations with weak π–π bonding, a dumbbell (DB) structure was proposed as a stable phase of group-IV films.33–36 Due to the fourfold-coordinated atoms included in this DB configuration, it is energetically more favorable than the LB configuration for some 2D group-IV materials,33,35,37–40 such as DB silicene and DB germanene. The plausibility of these DB configurations has also been evidenced by some experimental findings. For example, the image file: c5ra10712a-t1.tif silicene multi-layers grown on an Ag(111) substrate41–43 can be reproduced by stacking DB-based silicene derivatives.35,44 The 2 × 2 superstructure of germanene grown on an Al(111) substrate resembles DB germanene very well.45

More interestingly, the DB configuration of tin (DB stanene) has been predicted to be a QSH insulator with a large topologically nontrivial bulk band gap,33,46 especially when it is hydrogenated (DB stanane).46 These works enrich the database of group-IV QSH insulators, beyond the LB honeycomb lattices, such as silicene, germanene, δ-graphyne,47 graphene nanomesh,48 stanene,33 etc. However, the topological nontriviality is not retained in all the DB configurations of 2D group-IV materials. For example, the DB germanene proposed by Özçelik et al.,40 is a normal insulator. Moreover, group-IV elements can form abundant compounds with one another, e.g. silicon-carbide (SiC), silicon–germanium (SiGe), germanium–tin (SnGe), and so on. The honeycomb lattice of silicon carbide with a stoichiometry of SiC3 or Si3C has been predicted to be QSH insulators.25 However, the group-IV binary compounds with DB configurations have never been investigated so far. Whether they are QSH insulators remains unclear.

In this contribution, we reported our systematic study on the electronic structures of group-IV DB binary compounds (as shown in Fig. 1(a)) from first-principles calculations. We found that in contrast to DB stanene, none of them are QSH insulators. However, upon hydrogenation, germanium–tin (Sn6Ge4) can be tuned to a QSH insulator characterized by a topological invariant of Z2 = 1. The quantum phase transition was attributed to the band inversion of s–pxy of Sn atoms induced by surface hydrogenation. The nontrivial bulk band gap in hydrogenated DB Sn6Ge4 (Sn6Ge4H4) is about 235 meV, much larger than that of DB stanane46 and the thermal motion energy (∼26 meV) at room temperature. The phonon spectrum confirms the high stability and plausibility of DB Sn6Ge4H4. This work offers a promising candidate material for applications of QSH effect at high temperature.


image file: c5ra10712a-f1.tif
Fig. 1 (a) Schematic representations (top and side views are presented in up and middle panels) and the ELF profiles (down panel) of DB Sn6Ge4. (b) Phonon spectrum of DB Sn6Ge4 along the high-symmetric points in the BZ.

Methods

We preformed first-principles calculations within density-functional theory (EFT) using the plane wave basis Vienna Ab initio simulation package known as VASP code.49–51 The energy cutoff of the plane waves was set to 600 eV with an energy precision of 10−5 eV. A generalized gradient approximation (GGA) in the form proposed by Perdew, Burke, and Ernzerhof (PBE)52 was adopted to treat the electron exchange–correlation functional. The atomic positions were relaxed until the maximum force on each atom was less than 0.01 eV Å−1. The Brillouin zone (BZ) of the 2D materials was sampled by using an 11 × 11 × 1 gamma-centered Monkhorst–Pack grid. A vacuum space of up to 20 Å was applied to minimize the artificial interactions between neighboring slabs. SOC was included by a second variational procedure on a fully self-consistent basis. The phonon spectra were calculated using a supercell approach within the PHONON code.53

Results and discussion

The schematic representation of the DB configurations of group-IV binary compounds is shown in Fig. 1(a). Taking germanium–tin as an example, Sn and Ge have different coordination numbers, resulting in a stoichiometry of Sn6Ge4 (denoted as DB Sn6Ge4 hereafter). Sn is fourfold coordinated, similar to the case in stannane molecules (SnH4), while Ge is only threefold coordinated with a dangling bond as in the case of LB germanene. All the Sn atoms are on the middle plane sandwiched by two planes of Ge atoms. The buckling height of the 2D hexagonal structure is about 3.11 Å, shorter than that of DB stanene, 3.41 Å. The lattice constant (the length of the base vectors) is 8.83 Å. According to the electron localization function (ELF)54,55 profile shown in Fig. 1(a), despite the Ge–Sn covalent bonds, there are obvious regions of electron density in between the Ge atoms of the up- and down-planes, which are beneficial to the stabilization of the system. The length of the Ge–Sn bond is about 2.72 Å. The distance between the two Ge planes is about 3.11 Å, longer than that in germanene, 2.44 Å.

To evaluate the energetic stability of DB Sn6Ge4, we calculated the formation energy (Eform) with respect to LB germanene and stanene, according to the expression: Eform = (EtotalnGe × μGenSnμSn)/(nGe + nSn), where Etotal is the total energy per unit cell of DB Sn6Ge4, μGe and μSn are the chemical potentials of Ge and Sn atoms in the already-synthesized LB germanene and stanene,30 nGe are the respective numbers of the Ge and Sn atoms in one unit cell of DB Sn6Ge4. Our DFT calculations showed that DB Sn6Ge4 has a negative formation energy of about −0.14 eV per atom, suggesting the energetically favorability over LB germanene and stanene. The energetic stability of DB Sn6Ge4 is related to its unique coordination. The fourfold coordinated Sn and the interaction between the up- and down-planes of the threefold coordinated Ge atoms contribute to the stability of DB Sn6Ge4 over LB germanene and stanene composing exclusively of threefold coordinated Sn or Ge atoms.33–36 By ab initio phonon spectrum calculations, we further demonstrated the dynamic stability of DB Sn6Ge4, as shown in Fig. 1(b). No modes with imaginary frequencies were found in the spectrum and the structure was therefore proved to be dynamically stable.

The electronic band structures of DB Sn6Ge4 were then calculated. Without considering SOC, the band lines in the vicinity of the Fermi level, as indicated in Fig. 2(a), exhibit a direct band gap of 242 meV at the Γ point. When SOC was switched on, the direct-band-gap features were preserved, but the band gap was narrowed to 64 meV, as shown in Fig. 2(b). Parity analysis indicates that this is topological trivial band gap with a zero topological invariant (Z2 = 0). Pristine DB Sn6Ge4 is therefore a normal insulator, similar to the bulk counterpart. We also calculated the electronic band structures of other group-IV binary compounds with DB configurations, DB X6Y4 (X, Y = C, Si, Ge, and Sn), and found that none of them are QSH insulators (see ESI).


image file: c5ra10712a-f2.tif
Fig. 2 Band structures and their enlarged views (right panel) of DB Sn6Ge4 (a) without and (b) with SOC. The energy at the Fermi level was set to zero.

Ge atoms can be easily bonded with H atoms, for instance, in the case of germane (GeH4). We tried to passivate the dangling bonds of Ge atoms in DB Sn6Ge4 with H atoms. As shown in Fig. 3(a), all the Sn and Ge atoms in hydrogenated DB Sn6Ge4 (denoted as Sn6Ge4H4 hereafter) are fully-coordinated. The H–Ge distance in the optimized Sn6Ge4H4 is 1.56 Å, close to the value 1.55 Å in germane molecule. The distance between Ge planes is enlarged to 3.25 Å. Both the lattice constant (8.75 Å) and the Ge–Sn distance (2.70 Å) are shorter than the corresponding values of pristine DB Sn6Ge4 (8.83 Å and 2.72 Å), respectively. The binding energy of hydrogen atoms with DB Sn6Ge4 is about −2.17 eV, lower than that in DB stanane, −1.92 eV.46 These results confirmed that hydrogen atom can chemically bind to Ge atoms, in consistent with the ELF profile shown in Fig. 3(a). The Sn–Ge covalent bonds are retained in DB Sn6Ge4H4, while the interaction between the up- and down-planes of Ge atoms is weakened as indicated by ELF profiles and the enlarged spacing between Ge planes.


image file: c5ra10712a-f3.tif
Fig. 3 (a) Top (top panel) and side views (middle panel) and the ELF profiles (down panel) of a DB Sn6Ge4H4 unit cell. (b) Phonon spectrum of DB Sn6Ge4H4 along the high-symmetric points in the BZ.

To demonstrate the energetic stability of DB Sn6Ge4H4, we compared the energy of DB Sn6Ge4H4 with the sum of the energies of isolated DB Sn6Ge4 and atomic hydrogen atoms in hydrogen plasma environment.56,57 Our DFT calculations showed that the former is lower than the later by 0.62 eV per atom. The energetic favorability of DB Sn6Ge4H4 can also be demonstrated by a hypothetical reaction: 6SnH4 + 4GeH4 → Sn6Ge4H4 + 18H2. Our calculation indicates that it is an exothermal reaction with an energy release of 0.15 eV per atom. The phonon spectrum of DB Sn6Ge4H4 is plotted in Fig. 3(b). Obviously, there are no modes with imaginary frequencies, implying the dynamic stability of DB Sn6Ge4H4.

The electronic band structure of DB Sn6Ge4H4 differs significantly from that of pristine DB Sn6Ge4. Without considering SOC, DB Sn6Ge4H4 is a gapless semiconductor with the valence band maximum (VBM) and conduction band minimum (CBM) energetically degenerate at the Γ point, as shown in Fig. 4(a). The electronic states in the vicinity of the Fermi level originate from different atomic orbitals. The orbital-resolved electron density of states (DOS) showed that both VBM and CBM arise mainly from the px,y atomic orbitals of the Sn atoms, as shown in Fig. 4(b), in good consistent with the density profiles of the electron wavefunctions (WFs) shown in Fig. 4(d). For the valence band next to VBM (denoted as VBM-1), the WFs exhibit clear features of the s-orbital of Sn atoms and pz-orbital of Ge atoms. The degeneracy of the two px,y bands at the Γ point is a direct consequence of the C3v symmetry of the lattice. Such band alignment represents a common feature of a QSH insulator family, including DB stanane46 and the derivatives of germanene,31,32 which is closely related to topological nontriviality. Further studies on other group-IV binary compounds indicated that none of them exhibit such type of band alignment (see ESI). DB Sn6Ge4H4 stands out from the group-IV binary compounds as a candidate QSH insulator.


image file: c5ra10712a-f4.tif
Fig. 4 (a) The band structures without (top panel) and with SOC (down panel) of DB Sn6Ge4H4, with their enlarged views of band lines shown in the right panel. (b) Orbital-resolved electron density states of DB Sn6Ge4H4 without SOC. The DOS data were obtained by integrating the whole BZ. (c) and (d) are the isosurfaces of the electron wavefunctions of the VBM and VBM-1 at the Γ point. The isosurfaces of the CBM at the Γ point atom resemble those of the VBM and thus are not presented. The energy of the Fermi level was set to zero.

When SOC was switched on in our DFT calculations, a band gap of 275 meV is opened up at the Γ point of DB Sn6Ge4H4, as shown in the down panel of Fig. 4(a). The global indirect band gap is about 166 meV, slightly larger than that of DB stanane, 160 meV.46 Such a large bulk gap suggests that DB Sn6Ge4H4 is quite promising for achieving room-temperature QSH effect.

A comparison study of the orbital-resolved band structures of DB Sn6Ge4 and DB Sn6Ge4H4 was then conducted to determine their topological aspects. For convenience, we denoted the electronic states arising from s, px,y, pz atomic orbitals of Sn as |s±〉, |px,y±〉 and |pz±〉, in which the parity of the states are labeled by “±”. For pristine DB Sn6Ge4, no matter whether SOC is included or not, the energies of the electronic states near the Fermi level are successively in the same order: |px,y+〉, |pz〉 and |s〉, as shown in Fig. 5(a). But the order of |px,y+〉 and |s〉 is reversed in DB Sn6Ge4H4 without SOC, as shown Fig. 5(b), leading to a s–pxy band inversion. When SOC was included, the degeneracy of the partial occupied |px,y+〉 state in DB Sn6Ge4H4 was lifted, resulting in a large nontrivial band gap. This implies that surface hydrogenation induces the s–pxy band inversion and drives DB Sn6Ge4 to a QSH insulator, as visually revealed in Fig. 5(c). Such s–p-type band inversion mechanism59 also holds in other QSH insulators, such as HgTe quantum well7 and fluorinated stanene.31


image file: c5ra10712a-f5.tif
Fig. 5 Comparison between the orbital-resolved band structures of (a) DB Sn6Ge4 and (b) DB Sn6Ge4H4 in the vicinity of the Γ point obtained from DFT calculations without and with SOC. The superscript “+” and “−” denotes the parity of the wavefunctions. The contributions of different atomic orbitals are indicated by the dots in different colors and sizes. s-orbital of Sn (red), pxy-orbital of Sn (green), and pz-orbital of Sn (blue). (c) Schematic diagram of the electronic state evolution of DB Sn6Ge4 (left panel) and Sn6Ge4H4 (middle and right panel) near the Fermi level. The red line denotes the Fermi level, and the effects of (I) hydrogenation and (II) SOC on the alignment of electronic states are indicated. (d) Brillouin zone of DB Sn6Ge4 and DB Sn6Ge4H4. The signs (±) are associated with δi at the time-reversal invariant momenta. (b1 and b2) are the primitive reciprocal lattice vectors.

The topological nontriviality of DB Sn6Ge4H4 can be confirmed by a nonzero topological invariant Z2. Because both DB Sn6Ge4 and DB Sn6Ge4H4 have an inversion symmetry, the Z2 index can be calculated using the parity criteria proposed by Fu and Kane.58 According to this strategy, the Z2 index can be determined from the parities of the four time-reversal and parity invariant points at BZ. The four time-reversal invariant momenta of the two lattices occur at the Γ and three M points, as shown in Fig. 5(d). The Z2 invariant ν is defined by

image file: c5ra10712a-t2.tif
for 2N occupied bands. ξ2m(Γi) = ±1 is the parity eigenvalue of the 2m-th occupied energy band at the time-reversal invariant momentum Γi. Our DFT calculations showed that δi has the values of (+), (+), (+), (+) in DB Sn6Ge4 and (+), (−), (−), (−) in DB Sn6Ge4H4 at their four time-reversal invariant momenta, as indicated in Fig. 5(d). The topological invariants of pristine DB Sn6Ge4 and DB Sn6Ge4H4 are therefore Z2 = 0 and Z2 = 1 respectively, suggesting the quantum phase transition from a normal insulator (DB Sn6Ge4) to a QSH insulator (DB Sn6Ge4H4) induced by surface hydrogenation.

In view of the failure of PBE functional in determining electronic band gaps, we employed a hybrid functional in the form of Heyd–Scuseria–Ernzerhof (HSE)60 to recalculate the electronic band structure of DB Sn6Ge4H4. By incorporating a potion of exact exchange from Hartree–Fock (HF) theory, HSE functional can give band gaps comparable to experimental values. We found that the s–p band inversion and topological nontriviality of DB Sn6Ge4H4 are retained even under a tensile stain. The bulk band gaps opened due to SOC are about 235 meV (indirect) and 317 meV (direct band gap at the Γ), both of which are much larger than the PBE results.

Substrates are inevitable in the fabrication procedure of QSH-insulator-based devices. However, the coupling between QSH insulators with substrates may destroy their topological nontriviality.20,30 For instance, when stanane is grown on a semiconducting substrate, it becomes a normal insulator or a metal without topological nontriviality.30 To determine the influence of substrate on the topological nontriviality of DB Sn6Ge4H4, we built a superstructure of DB Sn6Ge4H4 on a (image file: c5ra10712a-t3.tif) h-BN substrate, as shown in Fig. 6(a). The lattice of DB Sn6Ge4H4 matches well with that of the h-BN substrate with a difference less than 0.6%. At the equilibrium state, the interlayer spacing between DB Sn6Ge4H4 and the substrate is about 2.29 Å, suggesting weak interaction between them. We calculated the electronic band structure of the superstructure and found that the band lines in the vicinity of the Fermi level are in accordance with the free-standing DB Sn6Ge4H4, as shown in Fig. 6(b) and (c). Both s–p band inversion and topological nontriviality are intact, except that the SOC bulk gap is slightly improved. The robust topological nontriviality of DB Sn6Ge4H4 against substrate effects is related to the origins of the bands near the Fermi level. These bands are mainly contributed by s and pxy atomic orbitals rather than pz orbital and thus are unlikely affected by substrates.


image file: c5ra10712a-f6.tif
Fig. 6 (a) Top and side views of a DB Sn6Ge4H4 grown on a h-BN substrate. Electronic band structures of DB Sn6Ge4H4 on h-BN substrate (b) without and (c) with SOC. The energy at the Fermi level was set to zero.

Conclusions

Using first-principles calculations, we demonstrated theoretically that the unique coordinates of the 2D germanium–tin compound with a dumbbell structure (DB Sn6Ge4) make it energetically more stable than low-buckled stanene and germanene. DB Sn6Ge4 is a normal insulator with a direct gap of 242 meV. However, upon surface hydrogenation (resulting in DB Sn6Ge4H4), it converts a QSH insulator characterized by a topological invariant of Z2 = 1. The quantum phase transition from a normal insulator to a QSH insulator is related to the s–p band inversion induced by surface hydrogenation. The bulk band gap opened due to SOC can be as large as 235 meV which is much larger than the thermal motion energy at room temperature (∼26 meV). The topological nontriviality is robust against substrate effects. The energetic stability and the phonon spectrum confirm the high stability and plausibility of DB Sn6Ge4H4. These results are beneficial for achieving long-desired room-temperature QSH effect and fabrication of high-speed spintronics devices.

Acknowledgements

This work is supported by the National Basic Research Program of China (No. 2012CB932302), the National Natural Science Foundation of China (No. 91221101, 21433006), the 111 project (No. B13029), the Taishan Scholar Program of Shandong, and the National Super Computing Centre in Jinan.

References

  1. X. L. Qi and S. C. Zhang, Topological insulators and superconductors, Rev. Mod. Phys., 2011, 83, 1057 CrossRef CAS.
  2. B. A. Bernevig and S. C. Zhang, Quantum spin Hall effect, Phys. Rev. Lett., 2006, 96, 106802 CrossRef.
  3. C. L. Kane and E. J. Mele, Quantum spin Hall effect in graphene, Phys. Rev. Lett., 2005, 95, 226801 CrossRef CAS.
  4. M. Z. Hasan and C. L. Kane, Colloquium: topological insulators, Rev. Mod. Phys., 2010, 82, 3045 CrossRef CAS.
  5. B. A. Bernevig, T. L. Hughes and S. C. Zhang, Quantum spin Hall effect and topological phase transition in HgTe quantum wells, Science, 2006, 314, 1757 CrossRef CAS.
  6. H. Weng, X. Dai and Z. Fang, Transition-metal pentatelluride ZrTe5 and HfTe5: a paradigm for large-gap quantum spin Hall insulators, Phys. Rev. X, 2014, 4, 011002 Search PubMed.
  7. C. L. Kane and E. J. Mele, Z2 topological order and the quantum spin Hall effect, Phys. Rev. Lett., 2005, 95, 146802 CrossRef CAS.
  8. C. Wu, B. A. Bernevig and S. C. Zhang, Helical liquid and the edge of quantum spin Hall systems, Phys. Rev. Lett., 2006, 96, 106401 CrossRef.
  9. C. Xu and J. Moore, Stability of the quantum spin Hall effect: effects of interactions, disorder, and Z2 topology, Phys. Rev. B: Condens. Matter Mater. Phys., 2006, 73, 045322 CrossRef.
  10. M. Konig, S. Wiedmann, C. Brune, A. Roth, H. Buhmann, L. W. Molenkamp, X. L. Qi and S. C. Zhang, Quantum spin Hall insulator state in HgTe quantum wells, Science, 2007, 318, 766 CrossRef PubMed.
  11. I. Knez, R. R. Du and G. Sullivan, Evidence for helical edge modes in inverted InAs/GaSb quantum wells, Phys. Rev. Lett., 2011, 107, 136603 CrossRef.
  12. Z. Liu, Z. F. Wang, J. W. Mei, Y. S. Wu and F. Liu, Flat Chern band in a two-dimensional organometallic framework, Phys. Rev. Lett., 2013, 110, 106804 CrossRef.
  13. L. Y. Li, X. M. Zhang, X. Chen and M. W. Zhao, Giant topological nontrivial band gaps in chloridized gallium bismuthide, Nano Lett., 2015, 15, 1296 CrossRef CAS PubMed.
  14. Z. F. Wang, N. Su and F. Liu, Prediction of a two-dimensional organic topological insulator, Nano Lett., 2013, 13, 2842 CrossRef CAS PubMed.
  15. Z. F. Wang, Z. Liu and F. Liu, Quantum anomalous Hall effect in 2D organic topological insulators, Phys. Rev. Lett., 2013, 110, 196801 CrossRef CAS.
  16. Z. Liu, C. X. Liu, Y. S. Wu, W. H. Duan, F. Liu and J. Wu, Stable nontrivial Z2 topology in ultrathin Bi (111) films: a first-principles study, Phys. Rev. Lett., 2011, 107, 136805 CrossRef.
  17. C. C. Liu, S. Guan, Z. G. Song, S. A. Yang, J. B. Yang and Y. Yao, Low-energy effective Hamiltonian for giant-gap quantum spin Hall insulators in honeycomb X-hydride/halide (X = N–Bi) monolayers, Phys. Rev. B: Condens. Matter Mater. Phys., 2014, 90, 085431 CrossRef.
  18. J. J. Zhou, W. Feng, C. C. Liu, S. Guan and Y. Yao, Large-gap quantum spin Hall insulator in single layer bismuth monobromide Bi4Br4, Nano Lett., 2014, 14, 4767 CrossRef CAS PubMed.
  19. F. C. Chuang, L. Z. Yao, Z. Q. Huang, Y. T. Liu, C. H. Hsu, T. Das, H. Lin and A. Bansil, Prediction of large-gap two-dimensional topological insulators consisting of bilayers of group III elements with Bi, Nano Lett., 2014, 14, 2505 CrossRef CAS PubMed.
  20. M. Zhou, W. Ming, Z. Liu, Z. Wang, P. Li and F. Liu, Epitaxial growth of large-gap quantum spin Hall insulator on semiconductor surface, Proc. Natl. Acad. Sci. U. S. A., 2014, 111, 14378 CrossRef CAS PubMed.
  21. M. W. Zhao, X. Chen, L. Y. Li and X. M. Zhang, Driving a GaAs film to a large-gap topological insulator by tensile strain, Sci. Rep., 2015, 5, 8441 CrossRef CAS PubMed.
  22. A. H. Castro Neto, N. M. R. Peres, K. S. Novoselov and A. K. Geim, The electronic properties of graphene, Rev. Mod. Phys., 2009, 81, 109 CrossRef CAS.
  23. J. C. Charlier and S. Roche, Electronic and transport properties of nanotubes, Rev. Mod. Phys., 2007, 79, 677 CrossRef CAS.
  24. R. Rurali, Colloquium: structural, electronic, and transport properties of silicon nanowires, Rev. Mod. Phys., 2010, 82, 427 CrossRef CAS.
  25. M. Zhao and R. Zhang, Two-dimensional topological insulators with binary honeycomb lattices: Si C 3 siligraphene and its analogs, Phys. Rev. B: Condens. Matter Mater. Phys., 2014, 89, 195427 CrossRef.
  26. L. Y. Li and M. W. Zhao, Structures, energetics, and electronic properties of multifarious stacking patterns for high-buckled and low-buckled silicene on the MoS2 substrate, J. Phys. Chem. C, 2014, 118, 19129 CAS.
  27. L. Y. Li and M. W. Zhao, First-principles identifications of superstructures of germanene on Ag (111) surface and h-BN substrate, Phys. Chem. Chem. Phys., 2013, 15, 16853 RSC.
  28. L. Y. Li, X. P. Wang, X. Y. Zhao and M. W. Zhao, Moiré superstructures of silicene on hexagonal boron nitride: a first-principles study, Phys. Lett. A, 2013, 377, 2628 CrossRef CAS PubMed.
  29. H. Zhou, M. Zhao, X. Zhang, W. Dong, X. Wang, H. Bu and A. Wang, First-principles prediction of a new Dirac-fermion material: silicon germanide monolayer, J. Phys.: Condens. Matter, 2013, 25, 395501 CrossRef PubMed.
  30. F. Zhu, W. Chen, Y. Xu, C. Gao, D. Guan, C. Liu, D. Qian, S. C. Zhang and J. Jia, Epitaxial growth of two-dimensional stanene, arXiv preprint arXiv:1506.01601, 2015.
  31. Y. Xu, B. Yan, H. J. Zhang, J. Wang, G. Xu, P. Tang, W. Duan and S. C. Zhang, Large-gap quantum spin Hall insulators in tin films, Phys. Rev. Lett., 2013, 111, 136804 CrossRef.
  32. C. Si, J. Liu, Y. Xu, J. Wu, B. L. Gu and W. Duan, Functionalized germanene as a prototype of large-gap two-dimensional topological insulators, Phys. Rev. B: Condens. Matter Mater. Phys., 2014, 89, 115429 CrossRef.
  33. P. Tang, P. Chen, W. Cao, H. Huang, S. Cahangirov, L. Xian, Y. Xu, S. C. Zhang, W. Duan and A. Rubio, Stable two-dimensional dumbbell stanene: a quantum spin Hall insulator, Phys. Rev. B: Condens. Matter Mater. Phys., 2014, 90, 121408 CrossRef.
  34. V. O. Özçelik and S. Ciraci, Local reconstructions of silicene induced by adatoms, J. Phys. Chem. C, 2013, 117, 26305 Search PubMed.
  35. S. Cahangirov, V. O. Özçelik, L. Xian, J. Avila, S. Cho, M. C. Asensio, S. Ciraci and A. Rubio, Atomic structure of the 3 × 3 phase of silicene on Ag (111), Phys. Rev. B: Condens. Matter Mater. Phys., 2014, 90, 035448 CrossRef.
  36. D. Kaltsas and L. Tsetseris, Stability and electronic properties of ultrathin films of silicon and germanium, Phys. Chem. Chem. Phys., 2013, 15, 9710 RSC.
  37. S. Cahangirov, M. Topsakal, E. Aktürk, H. Sahin and S. Ciraci, Two-and one-dimensional honeycomb structures of silicon and germanium, Phys. Rev. Lett., 2009, 102, 236804 CrossRef CAS.
  38. C. C. Liu, H. Jiang and Y. Yao, Low-energy effective Hamiltonian involving spin–orbit coupling in silicene and two-dimensional germanium and tin, Phys. Rev. B: Condens. Matter Mater. Phys., 2011, 84, 195430 CrossRef.
  39. C. C. Liu, W. Feng and Y. Yao, Quantum spin Hall effect in silicene and two-dimensional germanium, Phys. Rev. Lett., 2011, 107, 076802 CrossRef.
  40. V. O. Özçelik, E. Durgun and S. Ciraci, New phases of germanene, J. Phys. Chem. Lett., 2014, 5, 2694 CrossRef PubMed.
  41. P. de Padova, P. Vogt, A. Resta, J. Avila, I. Razado-Colambo, C. Quaresima, C. Ottaviani, B. Olivieri, T. Bruhn, T. Hirahara, T. Shirai, S. Hasegawa, M. C. Asensio and G. le Lay, Evidence of Dirac fermions in multilayer silicone, Appl. Phys. Lett., 2013, 102, 163106 CrossRef PubMed.
  42. P. Vogt, P. Capiod, M. Berthe, A. Resta, P. de Padova, T. Bruhn, G. le Lay and B. Grandidier, Synthesis and electrical conductivity of multilayer silicone, Appl. Phys. Lett., 2014, 104, 021602 CrossRef PubMed.
  43. B. Feng, Z. Ding, S. Meng, Y. Yao, X. He, P. Cheng, L. Chen and K. Wu, Evidence of silicene in honeycomb structures of silicon on Ag (111), Nano Lett., 2012, 12, 3507 CrossRef CAS PubMed.
  44. V. O. Özçelik, D. Kecik, E. Durgun and S. Ciraci, Adsorption of group-IV elements on graphene, silicene, germanene, stanene: dumbbell formation, J. Phys. Chem. C, 2015, 119, 845 Search PubMed.
  45. M. Derivaz, D. Dentel, R. Stephan, M. C. Hanf, A. Mehdaoui, P. Sonnet and C. Pirri, Continuous germanene layer on Al (111), Nano Lett., 2015, 15, 2510 CrossRef CAS PubMed.
  46. X. Chen, L. Li and M. Zhao, Dumbbell stanane: a large-gap quantum spin Hall insulator, Phys. Chem. Chem. Phys., 2015, 17, 16624–16629 RSC.
  47. M. Zhao, W. Dong and A. Wang, Two-dimensional carbon topological insulators superior to graphene, Sci. Rep., 2013, 3, 3532 Search PubMed.
  48. X. Zhang and M. Zhao, Prediction of quantum anomalous Hall effect on graphene nanomesh, RSC Adv., 2015, 5, 9875–9880 RSC.
  49. G. Kresse and J. Furthmüller, Efficient iterative schemes for ab initio total-energy calculations using a plane-wave basis set, Phys. Rev. B: Condens. Matter Mater. Phys., 1996, 54, 11169 CrossRef CAS.
  50. G. Kresse and J. Hafner, Ab initio molecular dynamics for open-shell transition metals, Phys. Rev. B: Condens. Matter Mater. Phys., 1993, 48, 13115 CrossRef CAS.
  51. G. Kresse and D. Joubert, From ultrasoft pseudopotentials to the projector augmented-wave method, Phys. Rev. B: Condens. Matter Mater. Phys., 1999, 59, 1758 CrossRef CAS.
  52. J. P. Perdew, K. Burke and M. Ernzerhof, Generalized gradient approximation made simple, Phys. Rev. Lett., 1996, 77, 3865 CrossRef CAS.
  53. K. Parlinski, Z. Q. Li and Y. Kawazoe, First-principles determination of the soft mode in cubic ZrO2, Phys. Rev. Lett., 1997, 78, 4063 CrossRef CAS.
  54. A. D. Becke and K. E. Edgecombe, A simple measure of electron localization in atomic and molecular systems, J. Chem. Phys., 1990, 92, 5397 CrossRef CAS PubMed.
  55. A. Savin, O. Jepsen, J. Flad, O. K. Andersen, H. Preuss and H. G. von Schnering, Electron localization in solid-state structures of the elements: the diamond structure, Angew. Chem., Int. Ed., 1992, 31, 187 CrossRef PubMed.
  56. Z. Sun, C. L. Pint, D. C. Marcano, C. Zhang, J. Yao, G. Ruan, Z. Yan, Y. Zhu and R. H. Hauge, Towards hybrid superlattices in graphene, Nat. Commun., 2011, 2, 559 CrossRef PubMed.
  57. M. Pumera and C. H. Wong, Graphane and hydrogenated graphene, Chem. Soc. Rev., 2013, 42, 5987 RSC.
  58. L. Fu and C. L. Kane, Topological insulators with inversion symmetry, Phys. Rev. B: Condens. Matter Mater. Phys., 2007, 76, 045302 CrossRef.
  59. H. Zhang and S. C. Zhang, Topological insulators from the perspective of first-principles calculations, Phys. Status Solidi RRL, 2013, 7, 72 CrossRef CAS.
  60. J. Heyd, G. Scuseria and M. Ernzerhof, Heyd–Scuseria–Ernzerhof hybrid functional for calculating the lattice dynamics of semiconductors, J. Chem. Phys., 2013, 118, 8207 CrossRef PubMed.

Footnote

Electronic supplementary information (ESI) available. See DOI: 10.1039/c5ra10712a

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