You are on page 1of 128

Progress in Materials Science 83 (2016) 24–151

Contents lists available at ScienceDirect

Progress in Materials Science


journal homepage: www.elsevier.com/locate/pmatsci

Rise of silicene: A competitive 2D material


Jijun Zhao a,⇑, Hongsheng Liu a, Zhiming Yu b, Ruge Quhe c,d, Si Zhou a, Yangyang Wang c,
Cheng Cheng Liu b, Hongxia Zhong c, Nannan Han a, Jing Lu c,e,⇑, Yugui Yao b,⇑, Kehui Wu f,e,⇑
a
Key Laboratory of Materials Modification by Laser, Ion and Electron Beams (Dalian University of Technology), Ministry of Education, Dalian 116024, China
b
School of Physics, Beijing Institute of Technology, Beijing 100081, China
c
State Key Laboratory of Mesoscopic Physics and Department of Physics, Peking University, Beijing 100871, China
d
State Key Laboratory of Information Photonics and Optical Communications & School of Science, Beijing University of Posts and Telecommunications, Beijing
100876, China
e
Collaborative Innovation Center of Quantum Matter, Beijing 100871, China
f
Beijing National Laboratory for Condensed Matter Physics and Institute of Physics, Chinese Academy of Sciences, Beijing 100190, China

a r t i c l e i n f o a b s t r a c t

Article history: Silicene, a silicon analogue of graphene, has attracted increasing attention during the past
Received 30 September 2015 few years. As early as in 1994, the possibility of stage corrugation in the Si analogs of
Received in revised form 5 April 2016 graphite had already been theoretically explored. But there were very few studies on
Accepted 7 April 2016
silicene until 2009, when silicene with a low buckled structure was confirmed to be
Available online 8 April 2016
dynamically stable by ab initio calculations. In spite of the low buckled geometry, silicene
shares most of the outstanding electronic properties of planar graphene (e.g., the ‘‘Dirac
Keyword:
cone”, high Fermi velocity and carrier mobility). Compared with graphene, silicene has
Silicene
several prominent advantages: (1) a much stronger spin–orbit coupling, which may lead
to a realization of quantum spin Hall effect in the experimentally accessible temperature,
(2) a better tunability of the band gap, which is necessary for an effective field effect
transistor (FET) operating at room temperature, (3) an easier valley polarization and more
suitability for valleytronics study. From 2012, monolayer silicene sheets of different super-
structures were successfully synthesized on various substrates, including Ag(1 1 1), Ir(1 1 1),
ZrB2(0 0 0 1), ZrC(1 1 1) and MoS2 surfaces. Multilayer silicene sheets have also been grown
on Ag(1 1 1) surface. The experimental successes have stimulated many efforts to explore
the intrinsic properties as well as potential device applications of silicene, including
quantum spin Hall effect, quantum anomalous Hall effect, quantum valley Hall effect,
superconductivity, band engineering, magnetism, thermoelectric effect, gas sensor, tunnel-
ing FET, spin filter, and spin FET, etc. Recently, a silicene FET has been fabricated, which
shows the expected ambipolar Dirac charge transport and paves the way towards
silicene-based nanoelectronics. This comprehensive review covers all the important theo-
retical and experimental advances on silicene to date, from the basic theory of intrinsic
properties, experimental synthesis and characterization, modulation of physical properties
by modifications, and finally to device explorations.
Ó 2016 Elsevier Ltd. All rights reserved.

⇑ Corresponding authors at: Dalian University of Technology, Dalian 116024, China (J. Zhao); Collaborative Innovation Center of Quantum Matter, Beijing
100871, China (J. Lu, K. Wu); School of Physics, Beijing Institute of Technology, Beijing 100081, China (Y. Yao).
E-mail addresses: zhaojj@dlut.edu.cn (J. Zhao), jinglu@pku.edu.cn (J. Lu), ygyao@bit.edu.cn (Y. Yao), khwu@aphy.iphy.ac.cn (K. Wu).

http://dx.doi.org/10.1016/j.pmatsci.2016.04.001
0079-6425/Ó 2016 Elsevier Ltd. All rights reserved.
J. Zhao et al. / Progress in Materials Science 83 (2016) 24–151 25

Contents

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26
2. Basic theory of silicene . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28
2.1. Stability of low-buckled structure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28
2.2. Mechanical properties. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30
2.2.1. Two-dimensional elastic constants . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30
2.2.2. Ultimate strength and fracture behavior. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31
2.3. Phonon dispersion and thermal properties . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32
2.3.1. Phonon dispersion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32
2.3.2. Mode-decomposed thermal transport . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33
2.3.3. Length dependence of thermal conductivity. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34
2.4. Electronic band structures and optical properties. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35
2.4.1. Band structures without spin–orbit coupling . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35
2.4.2. Optical properties . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36
2.5. Spin–orbit coupling and topological properties. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37
2.5.1. Spin–orbit coupling . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37
2.5.2. Quantum spin Hall effect . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 38
2.5.3. Extrinsic topological states . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39
2.6. Carrier mobility and electron–phonon interaction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 41
2.7. Superconductivity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 42
2.7.1. Phonon-mediated superconductivity. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 42
2.7.2. Unconventional superconductivity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43
3. Synthesis and characterization of silicene . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47
3.1. Silicene monolayer on Ag(1 1 1) substrate . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 48
3.1.1. T phase . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 48
3.1.2. Si(3  3)/Ag(4  4) phase . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 50
p p
3.1.3. 13  13R13.9° phase. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 53
p p p p
3.1.4. Si( 7  7)/Ag(2 3  2 3) phase . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 54
p p
3.1.5. ( 3  3)R30° phase of silicene . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 55
3.1.6. Interaction with Ag(1 1 1) and growth mechanism. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 57
3.1.7. Spectroscopic characterization and electronic structures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 60
3.2. Silicene monolayer on other substrates . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 66
3.3. Silicene nanoribbons . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 70
3.4. Multilayer silicene. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 73
4. Modification of silicene . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 78
4.1. Hydrogenation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 78
4.2. Halogenation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 83
4.3. Oxidization . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 85
4.4. Metal adsorption . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 88
4.5. Confinement effect . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 95
4.6. Defect effect. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 98
4.6.1. Stone–Wales defects. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 99
4.6.2. Vacancies. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 100
4.6.3. Nanomeshes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 101
4.6.4. Extended line defects . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 102
4.6.5. Adatoms . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 104
4.7. Strain effect . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 104
4.8. Semiconducting and insulating substrate effect . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 106
4.9. Multilayer heterostructures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 109
4.10. Substitutional doping and hybrid sheet. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 112
4.10.1. Substitutional doped silicene nanoribbon. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 112
4.10.2. Substitutional doped and hereto silicene sheets . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 112
5. Silicene electronics and spintronics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 114
5.1. Simulation of silicene FETs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 114
5.1.1. Effects of vertical electric field. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 115
5.1.2. Dual gated silicene FET . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 115
5.1.3. Alkali-adsorbed silicene FET . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 117
5.1.4. Silicene nanomesh FET . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 117
5.1.5. Tunneling FET . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 118
5.1.6. All-metallic FET. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 119
5.1.7. Silicene nanoribbon FET . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 119
5.2. Experimental progress in silicene FETs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 120
5.3. Silicene spintronics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 121
5.3.1. Spintronics in silicene nanoribbons. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 121
5.3.2. Spintronics in semihydrogenated functionalized silicene . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 125
26 J. Zhao et al. / Progress in Materials Science 83 (2016) 24–151

5.3.3. Silicene spintronic applications . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 126


6. Other applications of silicene . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 128
6.1. Thermoelectric devices . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 128
6.2. Chemical sensor. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 135
6.3. Hydrogen storage . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 136
6.4. Electrode material for Li battery . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 137
7. Summary and prospects . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 138
Acknowledgments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 140
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 140

1. Introduction

The coming out of graphene gives a verdict to whether single-atom-layer materials can exit. The Dirac cone of graphene
originated from the honeycomb lattice of carbon atoms endows graphene great prospect for future applications in nanoelec-
tronic devices [1] and stimulates significant efforts to explore other two-dimensional (2D) materials with honeycomb lattice
[2–5]. Among them, silicene, as the silicon analogue of graphene, exhibit many superior properties and continues to surprise
the community of low-dimensional materials.
As early as 1994, Takeda and Shiraishi investigated the Si and Ge analogues of graphite and discussed the planarity of the
2D single layers of Si and Ge using first-principles calculations [6]. In contrast to graphene where the A and B sublattices are
in exactly the same plane, they found that the two sublattices in silicene are relatively shifted in the direction perpendicular
to the atomic plane, forming a so-called low buckled structure. The stabilization of planar graphene was attributed to the
strong electron localization due to the shortest interatomic C–C distance. In 2007, silicene was named by Guzmán-Verri
and Lew Yan Voon [7], who developed a unified tight-binding (TB) Hamiltonian for Si-based nanostructures and showed that
silicene is a zero-gap semiconductor.
Until 2009, silicene was confirmed to be dynamically stable by Ciraci’s group via phonon dispersion calculations and
ab initio molecular dynamics (AIMD) simulations at T = 1000 K on the basis of density functional theory (DFT) [8]. Unlike
graphene, silicene is not absolutely planar, but with a buckled height of 0.44 Å. In spite of the low buckled geometry, the
band structure of silicene resembles that of graphene, that is, the p and p⁄ bands cross linearly at the Fermi level (EF), forming
the famous ‘‘Dirac cone” at the symmetric point K in the reciprocal space. Such band character indicates that the charge
carriers in silicene sheets behave like massless Dirac fermions with an estimated Fermi velocity of 106 m/s. Similar result
was obtained in an independent band structure calculation by Lebègue and Eriksson at the same year [9]. Moreover, Yao
et al. predicted a stronger spin–orbit coupling (SOC) effect in silicene than that in graphene due to heavier atom and buckled
geometry in silicene [10]. Therefore, more prominent quantum spin Hall effect (QSHE) can be anticipated in silicene in an
experimentally accessible temperature regime, which has greatly inspired the experimental synthesis of silicene.
Despite the outstanding electronic properties of silicene from the theoretical predictions, experimental synthesis of large-
scale silicene sheets remains a big challenge until 2012. Since there is no silicon allotrope with layered structure analogous
to the graphite in nature, the mechanical stripping method for fabricating graphene [1] and many other 2D materials [11]
does not work for silicene. Meanwhile, the chemical reduction method used for preparing graphene from graphene oxide
[12] cannot be applied to silicene either because that silicon oxide is a kind of hard insulator. The first experimental attempt
to prepare monolayer silicon sheets by chemical exfoliation of magnesium-doped calcium disilicide was reported by Nakano
et al. [13]. But the surface of the as-prepared silicon sheets was capped by oxygen atoms or probably other foreign atoms.
So far, the main method for synthesizing silicene is epitaxial growth on metal substrates. Before the large-scale mono-
layer sheets, finite-width silicene sheets (namely, nanoribbons) have been successfully fabricated on Ag(0 0 1) [14] and Ag
(1 1 0) [15–17] surfaces. Especially, the silicene nanoribbons on Ag(1 1 0) surface presented a magic width of 1.6 nm and
aligned parallel in a well distributed way [15]. The sp2-like hybridization of valence orbitals and graphene-like band disper-
sion in the Ag(1 1 0)-supported silicene nanoribbons were reported by reflection electron energy loss spectroscopy (REELS)
[16] and angle-resolved photoelectron spectra (ARPES) [17], respectively. Later, self-assembly of silicene nanoribbons
predominantly of 1.6 nm width have been obtained on the (2  1) reconstructed Au(1 1 0) surface [18].
In 2012, Vogt et al. [19] have epitaxially synthesized large-area silicene sheet on Ag(1 1 1) substrate and fulfilled the
theoretical prediction of silicene monolayer. The honeycomb lattice of silicene was evidenced by the Si 2p and Ag 4d core
level emission, the STM images, and the low energy electron diffraction (LEED) pattern. Based on the experimental observa-
tions and DFT calculations, they proposed a superstructure of (3  3) silicene lattice on (4  4) Ag(1 1 1) surface (referred as
4  4 superstructure thereafter). At the same year, several other groups [20–23] independently reported experimental
synthesis of monolayer silicene sheets on Ag(1 1 1) surface and observed a variety of superstructures, such as (4  4),
p p p p p p
( 13  13)R13.9°, (2 3  2 3)R30°, and 3  3. The superiority of Ag(1 1 1) as substrate for the growth of silicene
has been theoretically explained by Zhao et al. in term of the moderate and homogeneous interaction between silicene
and Ag surface as well as the low diffusion barrier of silicon atom on Ag(1 1 1) [24]. Beyond monolayer silicene, multilayer
silicene films were also fabricated on Ag(1 1 1) surface [20,25,26]. In addition to the Ag(1 1 1) substrate, monolayer silicene
sheets were successfully fabricated on metallic zirconium diboride (ZrB2) [27] thin films through epitaxial growth and on
J. Zhao et al. / Progress in Materials Science 83 (2016) 24–151 27

metallic ZrC(1 1 1) [28] and Ir(1 1 1) [29] surfaces by directly depositing silicon atoms. For the sake of device applications, it is
highly desirable to grow silicene on non-metallic substrate. Highly buckled silicene sheet was fabricated on a semiconduct-
ing MoS2 substrate by Molle et al. [30], opening the way towards non-metallic substrate growth of silicene.
Many theoretical efforts have been devoted to further exploring the other fundamental properties and possible applica-
tions of silicene, including quantum anomalous Hall effect, valleytronics, spintronics, superconductivity, band engineering,
field effect transistor (FET), tunneling FET, thermoelectric effect, gas sensor, etc. It is a quite natural idea to use silicene as a
channel material of a FET because of the compatibility with the mature silicon-based semiconductor technology and a high
carrier mobility, which is critical for a high speed device, and a better gate controllability due to the ultrathin 2D character.
Unfortunately, silicene suffers from the same difficulty as encountered by graphene—a zero band gap. In a conventional FET,
semiconducting channels with a sizeable band gap, preferably 0.4 eV or more are required to achieve a good switching
ability. Compared with graphene, silicene, however, has a better controllability of the band gap due to puckering in geometry
[31]. As pointed out by Lu et al. [31,32] and Drummond et al. [31,32], the first feasible way to achieve a band gap in silicene
without degrading its electronic properties is to apply a perpendicular electric field which breaks the symmetry of the two
sublattices of silicene. An alternative approach to open the band gap without degrading its electronic properties is adsorption
or intercalation of metal atoms [33–35]. Also benefited from its chemically active surface, the band gap of silicene can be
opened by chemical functionalization such as hydrogenation [36,37] and oxidization [38] that are accessible in experiments
[39,40]. Utilizing the quantum confinement effect, tunable band gaps up to about 0.4 eV were also predicted in silicene
nanoribbons by DFT calculations [8,41].
Despite the theoretical progress of the electronic devices based on silicene, it is very challenging to realize these devices
experimentally. This is because isolation of silicene from its template is difficult and the exposed silicene is generally unsta-
ble in air, leading it unfeasible to transfer silicene via the widely used wet transfer technique. Until recently, a novel grow-
transfer-fabrication technique has been developed and enables silicene FET, corroborating theoretical expectations regarding
its ambipolar Dirac charge transport [42]. Although the reported carrier mobility (100 cm2 V1 s1) and current on/off ratio
(10) are still low, there is large room to achieve higher mobility and better device performance. This breakthrough has
paved the way towards silicene-based 2D nanoelectronics, and it is generally believed that 2D nanoelectronics offer a
possible solution to enable Moore’s law in sub-10 nm region.
In silicene, the low-energy band structure connects different branches of physics, as the low-energy electrons behave like
massless Dirac fermions governed by the relativistic Dirac theory [10,43]. Because of the SOC effect, the pristine silicene is a
topological insulator which gives rise to the quantum spin Hall effect. Besides, silicene can host various kinds of topological
states by being applied external fields or by adsorption or intercalation of metal atoms. Notably, the electron–electron inter-
action in silicene is less screened compared to 3D materials, and may be responsible for many exotic physical phenomena,
especially the possible superconducting gap which has been observed in silicene epitaxially grown on Ag(1 1 1) surface [44].
Due to the fundamental and technological importance, silicene has attracted plenty of attentions in recent years. Since
2011, the number of publications on silicene increases dramatically, as displayed in Fig. 1. Although there have been some
excellent review articles on silicene [45–54], a comprehensive review covering all aspects of silicene, especially the most
recent progress of this fast-growing field is still imminent. Here we present a broad overview of current silicene research.
The rest of this article is organized as below. Section 2 summarizes the basic theory for the intrinsic properties of silicene,
including stability, elastic constants, ultimate strength, thermal conductivity, electronic band structures, spin–orbit coupling,
topological properties, carrier mobility, and superconductivity. The experimental synthesis of monolayer and multilayer
films of silicene on various substrates, the relevant spectroscopic characterizations and theoretical calculations, and some

Fig. 1. Trend of SCI-indexed publications concerning silicene (searched from Web of Science using the topic keyword: silicene).
28 J. Zhao et al. / Progress in Materials Science 83 (2016) 24–151

current debates are described in Section 3. In Section 4, we discuss various ways of modifying silicene, such as chemical
functionalization, metal decoration, patterning in nanoribbon or nanoflake, defect and strain engineering, interaction with
semiconducting and insulating substrates, stacking with other 2D materials, substitutional doping. The theoretical predic-
tions of silicene based electronic and spintronic devices, as well as recent experimental realization of silicene FET are
presented in Section 5. Section 6 covers the other applications of silicene, i.e., thermoelectric device, chemical sensor, hydro-
gen storage, and electrode material for Li battery. Finally, we summarize the essential aspects of silicene and give an outlook
of the future silicene research in Section 7.

2. Basic theory of silicene

2.1. Stability of low-buckled structure

The atomic structure of silicene is given in Fig. 2, showing the low-buckled honeycomb lattice. The calculated binding
energies of silicene and germanene as a function of the lattice constant [8] are presented in Fig. 3. Planar (PL),
low-buckled (LB), and high-buckled (HB) honeycomb structures are all local minima on the binding energy curves. The PL
honeycomb structures possess higher energy than the LB and HB counterparts and have imaginary phonon frequencies in
a large portion of the Brillouin zone (Fig. 3). Hence PL structures are not stable for silicene and germanene. Meanwhile,
the HB honeycomb structures of Si and Ge with a large buckling height of about 2 Å also have imaginary phonon frequencies.
Further optimization of HB structure on the (2  2) supercell results in an instability with a tendency towards clustering.
Therefore, the HB structures for Si and Ge do not correspond to real local minima on the Born–Oppenheimer surface either.
In contrast, the LB structures for Si and Ge exhibit positive values for the phonon dispersion curve in the entire Brillouin zone
(BZ) and thus are dynamically stable. The calculated Si–Si and Ge–Ge bond lengths for the LB structures are 2.2–2.4 Å and
2.4–2.5 Å, respectively.
Later, Roome and Carey [55] compared the phonon dispersion curves of silicene and germanene with PL, LB and HB
configurations and showed that the out-of-plane ZA and ZO phonon modes in the planar silicene and germanene contribute
to monolayer instability. Since the calculated electron-optical phonon coupling matrix elements in silicene and germanene
are 25 times smaller than in graphene, carrier relaxation via phonon scattering should be inhibited. Thus long momentum
relaxation lengths and high carrier mobilities can be expected in silicene and germanene based devices, making them
superior candidates for nanoelectronics.
The thermal stability of standalone silicene sheets were also explored by AIMD [8] and Monte Carlo [8,56] methods.
During AIMD simulation, silicene preserves its honeycomb lattice at temperature of 1000 K for 10 ps [8]. Using an empirical
Tersoff potential, the melting point of silicene from Monte Carlo simulations was estimated to be 1750 K, which is substan-
tially lower than that of bulk silicon (2200 K) described by the same potential [56]. Another MD simulation using reactive
force field (ReaxFF) predicted a critical temperature of Tc = 1500 K according to the Lindemann criteria of melting [57].
C, Si and Ge belong to the same group of the periodic table and thus have same number of valance electrons. However,
graphene prefers planar structure, while silicene and germanene tend to buckle. Such difference can be understood from the
orbital hybridization. For graphene, each C atom forms three r bonds with the adjacent C atoms via sp2 hybridization,

Fig. 2. Top view (upper panel) and side view (lower panel) of silicene lattice. The black rhombus represents primitive cell. The buckled height is depicted in
side view and denoted as h.
J. Zhao et al. / Progress in Materials Science 83 (2016) 24–151 29

leaving a perpendicular half occupied pz orbital. The coupling of pz orbital between adjacent atoms forms the p bond. Each
hexagonal ring in graphene lattice possesses one p bond, since there are two C atoms (i.e. two isolated electrons) in each
hexagonal ring. The planar structure of graphene arises from the short C–C distance and consequently the strong p bonding.
Compared to C, Si and Ge have larger atomic radii. Consequently, p bonds in silicene and germanene formed by coupling of
adjacent pz orbitals are much weaker than those in graphene due to the longer interatomic distance. Despite the weakened p
bonds, the stability of silicene and germanene is maintained by puckering induced dehybridization. As a result, the perpen-
dicular pz orbital combines with s orbital.
To explore the hybridization states in silicene and germanene, a comparative first-principles study on the orbital
hybridization in elementary 2D phases by carbon, silicon, and germanium was conducted by Wang [58]. Three dimensional
(3D) lonsdaleite crystals of C, Si and Ge were adopted as initial configurations. The 3D-to-2D transformation was simulated
by stretching the layer spacing d between two neighboring hexagonal basal planes gradually together with full relaxations of
the in-layer atomic position and the lattice parameter. In such way, the evolution of bonding state can be effectively
inspected from the variation of total energy E, as depicted in Fig. 4. For the carbon phase, there is an apparent turning point
from endothermic to exothermic in the E–d curve at d  2.38 Å, which is obviously due to the change of orbital hybridization
from sp3 to sp2. However, for the silicon and germanium phases, the total energy E keeps increasing with the increasing of d,
suggesting no such sp3-to-sp2 transition. In other words, the orbital hybridization remains sp3-like in silicene and germanene.
Taking hexasilabenzene (Si6H6) and other silicene-like clusters as model systems, Jose and Datta [59] explained the
microscopic origin of the buckling distortions in silicene by the strong coupling between the unoccupied molecular orbitals
and the occupied molecular orbitals, which leads to a pseudo-Jahn–Teller distortion (PJT). After buckling, the r-backbone is
stabilized and the p-backbone is destabilized. Opposite to the case of graphene, the former effect overwhelms the latter one
in silicene. Soto and co-workers [60] reexamine the PJT effect in silicene by performing DFT and time-dependent DFT
(TD-DFT) calculations of hexasilabenzene. By analyzing the vibronic instability of the ground state, they demonstrated that
a consistent PJT model should include not only the first excited state but also the next excited one, which is symmetry

Ó American Physical Society 2016

Fig. 3. Upper panel: binding energy versus hexagonal lattice constant of silicene and germanene. Black and dashed green curves of energy are calculated by
LDA using PAW potential and ultrasoft pseudopotentials, respectively. Planar and buckled geometries together with buckling distance D and lattice
constant of the hexagonal primitive unit cell, b are shown by inset. Lower panels: phonon dispersion curves obtained by force-constant and linear response
theory are presented by black and dashed green curves, respectively. Reproduced with permission from Ref. [8].
30 J. Zhao et al. / Progress in Materials Science 83 (2016) 24–151

Ó Royal Society of Chemistry 2016


Fig. 4. The E–d curves for lonsdaleite lattice from 3D to 2D structure evolution of carbon (upper), silicon (middle), and germanium (lower). The dashed lines
indicate the critical point for bond breaking. Reproduced with permission from Ref. [58].

Table 1
Second- and third-order elastic constants, in-plane Young’s modulus (E), and Poisson’s ratio m of monolayer silicene sheet from DFT calculations. All data are in
units of N/m except that m is dimensionless.

C11 C12 C111 C112 C222 E m


Zhang [63] 68.9 23.3 347.7 23.1 270.1 61.0 0.33
Peng [64] 71.3 23.2 397.6 14.1 318.9 63.8 0.325
Xu [65] 70.6 22.7 400.5 13.5 313.7 63.3 0.32

compatible coupled to the puckering mode. Recently, the same group further presented a PJT analysis of the puckering effect
in the decasilanaphthalene (Si10H8) molecule with two six-member ring [61]. Contrary to the hexasilabenzene with one
six-member ring, the multilevel PJT for the decasilanaphthalene molecule does not provide a satisfactory explanation for
puckering.

2.2. Mechanical properties

2.2.1. Two-dimensional elastic constants


Knowledge of the mechanical properties of silicene is crucial for designing silicene-based devices and understanding the
strength and reliability of these devices in practice. Though there is no mechanical experiment on silicene sheet yet, previous
measurement on the freestanding monolayer graphene revealed a large nonlinear elastic deformation during the tensile
strain until the intrinsic breaking strength (i.e., ultimate strength) is reached [62]. In other words, the nonlinear elastic prop-
erties are prominent in graphene, which might be a universal phenomenon for other 2D monolayer materials like silicene.
Using the homogeneous deformation method implemented in DFT calculations, both the second-order elastic constants
(SOEC) and third-order elastic constants (TOEC) of silicene have been computed to describe the linear and nonlinear elastic-
ity of silicene, respectively. The in-plane Young’s modulus (E), and Poisson’s ratio m can be further deduced from the second-
and third-order elastic constants by the following equations:

E ¼ ðC 211  C 222 Þ=C 11 ; t ¼ C 12 =C 11 : ð1Þ


J. Zhao et al. / Progress in Materials Science 83 (2016) 24–151 31

The theoretical results of silicene from several groups [63–65] are summarized in Table 1. One can see that different DFT
calculations yield nearly identical values of SOEC, whereas there are subtle differences in the TOEC, especially the C112.
Further analysis of the strain–energy curves indicated the third-order effects becomes dominant for strains larger than about
3.5% [63], showing the importance of nonlinear elasticity in silicene. Compared to graphene (E = 334 N/m, m = 0.18) [65],
silicene is much softer with only 20% of the in-plane stiffness and nearly twice of the Poisson’s ratio.

2.2.2. Ultimate strength and fracture behavior


In addition to the elastic constants which are the key parameters for continuum modeling of a material, the mechanical
behavior of silicene under tensile strains has been further simulated with various theoretical methods, including first-
principles DFT, semiempirical density functional tight-binding (DFTB), and empirical potentials (MEAM, Tersoff, ReaxFF).
As shown in Fig. 5, a typical stress–strain curve can be divided into three regions [66]: (i) the linear (harmonic) region, where
the Young’s modulus is defined; (ii) the nonlinear (anharmonic) region, where the nonlinear elasticity dominates; and (iii)
the plastic region, where the irreversible structural changes occur. The highest point of the strain–stress curve defines two
critical quantities: (i) the ultimate (or ideal) strength which is the maximum stress the system can withstand before ruptur-
ing, and (ii) the ultimate (or ideal) strain which is the critical strain at the moment of rupture.
The theoretical results of ultimate strengths and ultimate strains in current literature are summarized in Table 2. Despite
the discrepancies in the specific values due to different methodologies to describe the total energy and force, most

Ó American Institute of Physics 2016

Fig. 5. The stress–strain curves of silicene at different temperatures under tension along (a) armchair and (b) zigzag directions. The effect of temperature on
(c) the fracture strength and (d) fracture strain of silicene. Reproduced with permission from Ref. [71].

Table 2
Ultimate in-plane strengths and ultimate strains of monolayer silicene under armchair (AC), zigzag (ZZ) and equiaxial (EQ) tensions from DFT calculations and
MD simulations with classical potentials.

Ultimate strength (N/m) Ultimate strain


AC ZZ EQ AC ZZ EQ
Zhao (DFT, 0 K) [69] 7.07 5.66 – 0.18 0.14 –
Yang (DFT, 0 K) [67] 7.59 5.26 6.76 0.17 0.136 0.16
Peng (DFT, 0 K) [64] 6.0 5.9 6.2 0.17 0.21 0.17
Roman (ReaxFF, 10 K) [70] 4.78 5.85 – 0.089 0.18 –
Pei (MEAM, 300 K) [71] 3.85a 3.94a – 0.18 0.21 –
a
The in-plane strengths were calculated by assuming a thickness of 0.313 nm for silicene sheet.
32 J. Zhao et al. / Progress in Materials Science 83 (2016) 24–151

theoretical simulations show consistent results. First of all, comparison between different tension orientations (armchair vs.
zigzag) shows that the ultimate strain is anisotropic, although the Young’s moduli are virtually isotropic for the silicene
membranes [66]. All DFT calculations indicated that the ideal strength of silicene along armchair (AC) orientation is higher
than that along zigzag (ZZ). Under zigzag uniaxial tension, a phase transition of silicene from the original low-buckle struc-
ture to completely planar one occurs at a critical strain of 0.16 [67] or 0.20 [68], beyond which the planar hexagonal silicene
is mechanically stable from phonon dispersion analysis. The failure mechanism of silicene under AC and ZZ uniaxial tensions
is elastic instability and phonon instability occurring behind the elastic instability, while the failure mechanism of silicene
under equiaxial tension is only attributed to elastic instability [67]. More interestingly, the strain engineering of Poisson’s
ratio of silicene exhibits strong anisotropy: it increases when stretched in the ZZ direction and could be anomalously greater
than 0.5; but it decreases under AC uniaxial strain [66].
In addition to DFT calculations at zero temperature, empirical MD simulations are able to reveal the temperature depen-
dence of mechanical strength and present the detailed fracture behavior. At higher temperature, both the fracture strength
and fracture strain reduce significantly (Fig. 5) probably due to the following two factors [66,71]: (i) the Si–Si bond lengths
experience larger fluctuations at higher temperatures, making some bonds easy to reach the critical bond length and break;
(ii) the initial Si–Si bonds are longer at higher temperatures and thus are more easily to reach the breaking point during
tensile strain [71].
The stress–strain curves in Fig. 5 show that no apparent plastic deformation takes place before fracture. Once reaching the
fracture strength, the stresses drop abruptly. Therefore, silicene exhibits brittle fracture under tension deformation, which
was also observed in MD simulation of fracture processes, as illustrated in Fig. 6. Although the tension deformation is
uniformly applied over the whole silicene area, the atomic stresses are not uniformly distributed due to thermal fluctuation
at finite temperature. At a strain of 0.182 (Fig. 6a), a small crack is nucleated by direct breaking of a Si–Si bond, which reaches
its critical bond length. Afterwards (Fig. 6b and c), the crack starts to propagate spontaneously, resulting in cleavage failure
mode without plasticity.

2.3. Phonon dispersion and thermal properties

The thermal properties (specially, thermal conductivity) of silicene play a key role in the reliability and performance of
silicene-integrated electronic and optoelectronic devices [72]. Hence, the thermal conductivity of silicene has received
increasing attentions in recent years. It is well known that graphene exhibits an ultrahigh thermal conductivity of 2500–
5000 W/mK at room temperature [73,74]. Due to the similarities and differences of the lattice structures between silicene
and graphene, it is of great interest to compare their phonon transport mechanisms and to elucidate the effects of structural
buckling on the thermal conductivity of silicene. Here we discuss the phonon transport properties of pristine silicene from
the following three aspects: (i) phonon dispersion, (ii) mode-decomposed thermal transport, (iii) length dependence of
thermal conductivity. The extrinsic effects of confinement, strain and doping on the thermal conductivity of silicene will
be discussed in Section 4.

2.3.1. Phonon dispersion


The phonon dispersions of graphene and silicene are shown in Fig. 7. Both graphene and silicene have six phonon
branches: flexural acoustic (ZA), transverse acoustic (TA), and longitudinal acoustic branches (LA), flexural optical (ZO),
transverse optical (TO), and longitudinal optical branches (LO). Unlike graphene whose flexural mode is purely out-of-
plane vibration, the flexural mode of silicene contains small components along the in-plane directions [72,75]. This is
because the buckled structure of silicene breaks the reflectional symmetry over the xy plane, leading to the coupling of
out-of-plane and in-plane vibration modes. In addition, the ZA branch of silicene contains a linear component near C point
due to its buckled structure, yielding a small but non-zero sound velocity of 1010 m/s, while for graphene it follows a quad-
ratic trend near C point [72]. Compared to graphene, the TA and LA branches of silicene are less dispersive, owing to the
weaker strength of Si–Si bonds than C–C bonds. The average sound velocities of TA and LA branches for graphene are
13.6 km/s and 21.3 m/s, respectively [73], and those for silicene are 5.4 km/s and 8.8 m/s, respectively [76]. The frequencies
Ó American Institute of
Physics 2016

Fig. 6. The stress distribution, deformation, and failure process of silicene under tension. (a) Crack nucleation through local bond breaking; (b and c) crack
growth; (d) fracture of structure. Reproduced with permission from Ref. [71].
J. Zhao et al. / Progress in Materials Science 83 (2016) 24–151 33

Ó American Institute of Physics 2016


Fig. 7. Phonon dispersions of (a) graphene, and (b) silicene by first-principles calculations. Reproduced with permission from Ref. [72].

of the LO (or TO) modes at the C point are 1630 cm1 and 570 cm1 for graphene and silicene, respectively, corresponding to
the G peak in Raman signal [72].
Beyond the harmonic approximation, the thermal expansion, Grüneisen constant (c) and temperature-dependent
stiffness of silicene have been investigated by Zeng’s group using DFT calculations with a self-consistent quasiharmonic
approximation [77]. They found that bond bending (stretching) negatively (positively) contributes to c, which leads to
the negative acoustic (positive optical) c. The excitation of the negative-c modes further results in negative thermal
expansion coefficients for silicene within the temperature region of 0–600 K.

2.3.2. Mode-decomposed thermal transport


So far, substantial efforts have been made to calculate the thermal conductivity of silicene, based on classical molecular
dynamics simulations by using Tersoff potential [78,79] and Stillinger–Weber (SW) potential [80], or by using first-principles
DFT [72,75]. The intrinsic thermal conductivity of silicene at room temperature was found to be 3–65 W/mK [72,75,78–80],
depending on the computational methods employed. Nevertheless, it is much lower than those of graphene and bulk silicon
(about 150 W/mK at room temperature) [80].
The difference of the thermal conductivity between graphene and silicene can be understood by considering the phonon
Boltzmann transport equation (BTE) [80]:
X
j¼ cp m2p sp ; ð2Þ
p

where j is the lattice thermal conductivity, c is the heat capacity, v is the group velocity, s is the phonon relaxation time, and
the sum is taken over all phonon modes p of the system. Since the optical branches have smaller group velocities relative to
the acoustic branches, the thermal conductivity is dominated by the three acoustic modes.
The intrinsic phonon transport is limited by the phonon–phonon scattering due to the anharmonicity of lattice vibrations,
which is described as the Umklapp (U) process [73]. The thermal conductivity is extrinsic when phonon-boundary and
phonon-defect scattering become prominent [74]. Fig. 8 presents the phonon U-process scattering rates of individual acous-
tic modes for graphene and silicene. Overall speaking, silicene has larger phonon scattering rates than graphene. The larger
scattering rate, the more likely a phonon is to be scattered (and thus the shorter lifetime).
Ó American Institute of Physics 2016

Fig. 8. Scattering rates of acoustic phonon modes for (a) graphene, and (b) silicene along U–M direction at T = 300 K. Reproduced with permission from Ref.
[72].
34 J. Zhao et al. / Progress in Materials Science 83 (2016) 24–151

Moreover, the ZA mode of graphene has a much smaller scattering rate than the TA and LA modes, and it contributes to
about 80% of thermal conductivity, as shown in (Fig. 9a) [72]. For silicene, the scattering rate of ZA mode is comparable to
those of the other two acoustic modes. According to BTE, the group velocity of the ZA branch for silicene is about one order of
magnitude lower than those of the TA and LA branches, thus the contribution of the ZA mode to thermal conductivity is less
than 10% in silicene (Fig. 9b) [72,75]. Therefore, the low thermal conductivity of silicene can be attributed to two factors: (i)
the weaker Si–Si bonding strength gives less dispersive acoustic branches compared to graphene, leading to lower phonon
group velocities and relaxation time; (ii) the structural buckling of silicene results in a coupling between the out-of-plane
and in-plane vibration modes, hindering the heat conduction carried by the ZA mode.

2.3.3. Length dependence of thermal conductivity


The size effect on the intrinsic thermal conductivity of 2D materials is an important and unsettled issue. It gives informa-
tion on the phonon coherent length of a system. For graphene, theoretical studies suggested that its thermal conductivity
diverges with the transport length [81,82]. Accordingly, experiments found very long but finite phonon mean free path in
graphene due to impurities in the graphene samples [83,84].
For silicene, a few numerical simulations have been carried out to explore the length dependence of thermal conductivity;
but the results are contradictory [72,78–80]. The non-equilibrium molecular dynamics (NEMD) simulations using SW
potentials showed that the thermal conductivity of silicene at room temperature increases rapidly with the sample length
below 300 nm and reaches a plateau thereafter (Fig. 10a). The diffusive thermal conductivity of silicene is estimated to be
8.64 W/mK, and the phonon mean free path is 23.76 nm [80]. Similar results were reported based on the NEMD method,
showing that the thermal conductivity of silicene sheet attains convergence with transport length beyond about 60 nm
[78,79]. On the contrary, the thermal conductivity was found to be insensitive to the transport length, from the equilibrium
molecular dynamics (EMD) simulations and the anharmonic lattice dynamics (ALD) calculations using SW potentials

Ó American Institute of Physics 2016


Fig. 9. Relative thermal conductivity of difference phonon modes for (a) graphene, and (b) silicene, with a sample length of 10 lm as a function of
temperature. Reproduced with permission from Ref. [72].
Ó American Institute of Physics 2016

Fig. 10. Thermal conductivity of silicene as a function of sample size by (a) NEMD [80] and (b) DFT calculations. Reproduced with permission from Ref.
[72].
J. Zhao et al. / Progress in Materials Science 83 (2016) 24–151 35

Ó American Institute of Physics 2016


Fig. 11. Electronic band structures (left panel) and density of states (right panel) of silicene with low-buckled configuration calculated without inclusion of
spin–orbit coupling. Reproduced with permission from Ref. [92].

Table 3
Fermi velocity (vF) of low-buckled silicene and planar graphene from band structures calculations using different methods.

Method Silicene Graphene


LDA [85] PW91 [86,90] HSE06 [91] GW [85] PW91 [86,90] HSE06 [91]
vF (m/s) 5.4  105 5.32  105 6.5  105 7.4  105 8.29  105 10.1  105

(Fig. 10a) [80]. A combination of first-principles based ALD method and the BTE formalism yields divergent thermal conduc-
tivity of silicene with the transport length L up to 10 lm (Fig. 10b) [72]. In addition, it was shown that the contributions from
optical modes and flexural acoustic mode become constant when L > 1 lm. The in-plane acoustic phonons are responsible
for the continuous increase of the thermal conductivity with respect to the sample length. The ambiguous results on the
length dependence of thermal conductivity of silicene are due to the different simulation techniques and approximations
employed, such as the empirical potentials used in MD simulations, periodic conditions applied, and limited size of simula-
tion supercell.
The accumulated thermal conductivity as a function of phonon mean free path (MFP) can provide insights into the size effect.
It was shown that at room temperature more than 50% of the overall thermal conductivity is contributed by the phonons with
mean free paths less than 10 nm. This is completely different from bulk silicon, in which nearly 50% of the thermal conductivity
is attributed by the phonons with mean free paths longer than 1 lm [80]. These results imply a shorter coherent phonon length
of silicene than that of bulk silicon. However, MFP is not the only factor that affects thermal conductivity, and the wavelength is
also a concern. Phonons having longer wavelengths may not contribute to heat conduction by carrying heat themselves, but
they scatter with other phonons [80]. To include the long-wavelength phonons is important for studying the convergence of
thermal conductivity with transport length. Continuous efforts have to be made in both experiment and theory to clarify the
size effect on the intrinsic phonon transport properties of silicene as well as other 2D materials.

2.4. Electronic band structures and optical properties

2.4.1. Band structures without spin–orbit coupling


From DFT calculations (without inclusion of spin–orbit coupling) of electronic band structures [7,8,41,85–90], silicene is a
semimetal (or zero-gap semiconductor) with valence band maximum (VBM) and conduction band minimum (CBM) touched
at the K point in the first BZ, as displayed in Fig. 11. Similar to graphene, the p and p⁄ bands of silicene are crossed linearly at
the Fermi level, resulting in a massless Dirac fermion character of the charge carriers. Note that standard DFT calculations of
silicene without spin–orbit coupling always leads to a zero-gap semiconductor (or semimetal); while spin–orbit interaction
opens a fundamental band gap of magnitude of about 1.55 meV at the Dirac points for silicene [10], which will be discussed
in Section 2.5.
From the linear dispersion of characteristic p-bands near Fermi level, the Fermi velocity (vF) of carriers of silicene can be
estimated from a linear fitting of wave vector k by neglecting the second and higher order terms:
1 dE
vF    ð3Þ
h dk
36 J. Zhao et al. / Progress in Materials Science 83 (2016) 24–151

As summarized in Table 3, vF is about 5  105 m/s from conventional DFT calculations, which is moderately lower than
that for graphene (8.29  105 m/s) [90] but still one order of magnitude higher than that of Si(1 1 1) layer (104 m/s) [7].
The reduced vF of silicene with regard to graphene is attributed to the fact that electron tunneling between Si atoms is less
favorable in silicene than in graphene because of the weaker p interaction [7]. Inclusion of self-energy corrections using the
single-shot G0W0 approximation results in 37% enhancement of vF from the DFT value because of the depressed screening
effect [85]. Similarly, hybrid HSE06 functional with partial inclusion of exact exchange interaction leads to an increase of vF
by 22% [87].
It is also noteworthy that the buckling geometry has little effect on the Fermi velocity, for instance, vF = 5.6  105 m/s for
planar silicene and vF = 5.4  105 m/s for buckled one from the same LDA calculations [85]. Furthermore, Rojas-Cuervo et al.
[89] found that the Dirac cones of silicene in the vicinity of the K points display two kinds of asymmetries: (i) deviation from
circular shape and exhibiting the symmetry of an equilateral triangle (D3), (ii) the hole Dirac cones being shorter and more
asymmetric than the electron Dirac cones. As a consequence, the Fermi velocities for electrons and holes are direction depen-
dent, that is, the vF for electron (hole) carrier along the K–C direction is higher than that along K–M by 61.4% (57.5%). Similar
asymmetry was found in graphene, where the relative variation of electron (hole) Fermi velocity between K–M direction and
K–C direction is 36.6% (28.7%) [89]. Obviously, the asymmetry of the Dirac cone of silicene is more pronounced than that of
graphene.

2.4.2. Optical properties


From the electronic band structures, the electron excitation behavior and optical properties of silicene can be further
derived from first-principles calculations [86–88,90,91,93–95]. Matthes et al. [86,90,91,94] carried out systematical calcula-
tions of the optical properties of silicene and its counterparts, i.e., graphene, germanene, tinene (or stanene). In their studies,
the optical absorbance of silicene was derived from ab initio calculation of the complex dielectric function for optical inter-
band transitions in the framework of the independent-quasiparticle approximation, whereas the many-body effects are
incorporated in the framework of a non-local screened hybrid HSE06 functional. As depicted in Fig. 12, due to strong
inter-band transition at the M point in the 2D BZ, the saddle point in the difference ec(k)  ev(k) of the lowest conduction
band (p⁄-like) and the highest valence band (p-like) give rise to a pronounced peak (M1) at 4.0 eV (graphene), 1.6 eV

Fig. 12. (Upper panel) Spectral absorbance (in units of pa) for graphene (black solid line), silicene (red dashed line), and germanene (blue dotted line).
Reproduced with permission from Ref. [86]. Copyright 2016, American Physical Society. (Lower panel) Optical conductivity in units of the ac conductivity r0
of silicene. Real part: black line, imaginary part: red line. Reproduced with permission from Ref. [94]. Copyright 2016, IOP Publishing.
J. Zhao et al. / Progress in Materials Science 83 (2016) 24–151 37

(silicene), or 1.7 eV (germanene), respectively. The most pronounced peak in the absorption spectrum of silicene at 3.9 eV is
mainly related to the M2 maximum of the interband transition energy at the C–M line. After partly including the exact
exchange by HSE06 functional, the optical absorption peaks are blue-shifted, while the influence on the low-frequency
absorbance remains unchanged. For the optical conductivity of silicene shown in Fig. 12, the most important intensity
variations with the photon energy can be simply described by two damped harmonic oscillators which may be traced back
to some van Hove singularities in the 2D interband structure. In the real part of optical conductivity, the first peak near 2.0 eV
is a consequence of 2D saddle points in the p⁄ ? p interband structure located at the six M points at the BZ boundary, while
the second peaks near 5.0 eV are related to the r ? r⁄ transitions mainly at the C point of the 2D BZ.
In an independent study, Wei and co-workers [87,88] investigated the many-body effects in silicene based on the Green’s
function perturbation theory, i.e., GW + Bethe–Salpeter equation. They observed optical resonances in silicene due to
excitonic effects: the p ? p⁄ excitonic resonance at 1.23 eV is contributed by the characteristic dispersion of Dirac fermions,
while the one at 3.75 eV is due to the r ? p⁄ transitions.
In addition to monolayer silicene, Mohan et al. [93] computed the dielectric function and electron energy loss (EEL) func-
tion of bilayer silicene in AA and AB stacking modes. In low frequency region (0–5 eV), there are no electron excitations for
out-of-plane polarization and for in-plane polarization, and the plasmons are found at 2.16 eV for monolayer and at 1.48 eV
for AB-stacked bilayer silicene, respectively. For in-plane polarization, the EEL spectra show p (p + r) plasmonic features at
2.16 eV (7.60 eV), 1.48 eV (10.54 eV) and (10.5 eV) in monolayer, AB-stacked and AA-stacked bilayer silicene, respectively.
Multiple p + r surface plasmons of about 10 eV are also seen in out-of-plane polarization of monolayer and bilayer silicene
sheets.

2.5. Spin–orbit coupling and topological properties

Topological order is a classification paradigm beyond the picture of Landau’s spontaneous symmetry breaking [96,97],
and the states with the topological order can possess some distinct properties. For example, 2D topological insulator (TI),
unlike the ordinary band insulator (BI), exhibits not only a bulk gap but also metallic edge states, which are topologically
protected and immune to the disorders and defects [98–100].
Generally, the band topology of an insulator is characterized by Z2 topological invariant and Chern numbers. Z2 topolog-
ical invariant is an obstruction to make the wave functions smoothly defined over half of the entire BZ under a certain gauge
with the time reversal (TR) constraint [101–103]. Z2 = 0, 1 correspond to band insulator and topological insulator, respec-
tively [104–106]. On the other hand, Chern number is related to the Berry curvature and well defined regardless of the
existence of the TR symmetry. But when the TR symmetry is present, the total Chern number is always zero. Thus, other
kinds of Chern number, such as spin-Chern number, have to be introduced [107–110]. Furthermore, in materials with
honeycomb structure, the valley-Chern number can be also introduced naturally.
Based on the edge-bulk correspondence, there is also another way to explore the band topology of an insulator by ana-
lyzing the edge energy spectrum. The edges separate the system and the topologically trivial vacuum. Hence, if the system is
a topological insulator, gapless edge modes will emerge on its edge because the topological number cannot change without a
gap closing.
In the past decade, one inspiring and pioneering discovery is that the nontrivial properties of a system may stem from the
spin–orbit interaction. Graphene is the first proposed SOC-induced topological insulator – quantum spin Hall insulator
[111,112], possessing nonzero Z2 invariant and spin Chern numbers. But subsequent works showed that the SOC of graphene
is extremely weak and QSHE in graphene can occur only at unrealistically low temperature [113].
Silicene, instead of graphene, was demonstrated to be a good candidate to realize QSHE, because it exhibits considerably
large SOC gap [10]. The pristine germanium and stanene are quantum spin Hall insulators with large SOC gap too, as also
pointed out by Yao et al. [10]. Besides, since the electronic properties of silicene can be effectively tuned by external fields
[114–124], such as electric field, exchange field and electromagnetic field, various kinds of topological states can be realized
in silicene.

2.5.1. Spin–orbit coupling


In general, SOC in the Pauli equation reads:

h
Hso ¼  ðF  pÞ  r; ð4Þ
4m20 c2

where F is the force, p is the momentum,  h is the Plank constant, m0 is the mass of free electron, c is the light speed, and r is
the Pauli matrices. For silicene, the nearest-neighbor SOC is zero, while the next-nearest-neighbor SOC is nonzero and can be
divided into two parts from in-plane and out-of-plane contributions, corresponding to intrinsic SOC and intrinsic Rashba
SOC, respectively [43], as shown in Fig. 13.
Although the geometries of graphene and silicene are similar, the manifestations of SOC effect in them show some distinct
differences. In graphene, since the two sublattices are coplanar (h = 90°), the intrinsic effective first-order SOC and intrinsic
Rashba SOC vanish [43], resulting in a rather tiny SOC gap from the second-order SOC [113]. In silicene, however, due to the
38 J. Zhao et al. / Progress in Materials Science 83 (2016) 24–151

Ó American Physical Society 2016


Fig. 13. The atomic intrinsic SOC and intrinsic Rashba SOC. (a) The nearest-neighbor force F1 vanishes, while the next-nearest-neighbor force F|| is nonzero
in the horizontal plane. (b) The next-nearest-neighbor nonzero force F ?A equals negative F B? in the perpendicular direction. (c) h is defined as the angle
between the bond and the z direction normal to the plane. Reproduced with permission from Ref. [43].

Ó American Physical Society 2016


Fig. 14. (a) The variation of the SOC gap at the Dirac point with the angle h for silicene. Black line marks the total gap. Red or gray line and dotted line denote
gaps opened by the first- and second-order SOC, respectively. (b and c) Sketches of the intrinsic effective first- and second-order spin–orbit interactions.
Reproduced with permission from Ref. [43].

low-buckled geometry (h > 90°), the intrinsic effective SOC is a first-order process and the intrinsic Rashba SOC emerges [43].
Moreover, the gap opened by SOC increases with h [10,43], as shown in Fig. 14.

2.5.2. Quantum spin Hall effect


In pristine silicene, the SOC opens a band gap at the Dirac points, and gives rise to the QSHE, which has been confirmed by
the calculation of Z2 topological invariant with first-principles calculations [10]. On the other hand, the topological properties
of silicene can be also explored by the following TB model. The TB Hamiltonian of pristine silicene reads [43]:
X tiso X 2t X
H ¼ t cyia cja þ i pffiffiffi tij cyia razb cjb  i iRso li cyia ðr  dij Þazb cjb ; ð5Þ
hi;jia 3 3 hhi;jiiab
3 hhi;jiiab

where cyia ðcia Þ is the electron creation (annihilation) operator with spin a at site i, r is the Pauli matrix, dij represents a unit
vector connecting sites i and j, hi, ji (hhi; jii) stands for nearest (next-nearest) neighbor sites. Generally, the parameters of the
TB model can be fitted from the first-principles calculations of band structures of pristine silicene [43]. In Eq. (5), the first
term is the usual nearest-neighbor hopping with t = 1.08 eV. The second term describes the intrinsic SOC with tiso = 3.9 meV,
and tij = 1 (1) if the next-nearest-neighbor hopping is counterclockwise (clockwise) with respect to the positive z axis. The
last term denotes the intrinsic Rashba SOC with tiRso = 0.7 meV, and li = 1 (1) for i at A(B) sublattice.
The low-energy band structure of silicene situates in the vicinity of valley points (K/K0 ), and the low-energy Hamiltonian
is written as:

Hg ¼ htF ðgkx sx þ ky sy Þ þ gtiso sz sz þ tiRso sz ðky asx  kx asy Þ; ð6Þ

where g = 1 (1) represents valley index (K/K0 ). s and s are Pauli matrices representing the A/B sublattice and spin, respec-
pffiffi
tively. tF ¼ 23 at ¼ 5:5  105 m=s is the Fermi velocity with the lattice constant a = 3.86 Å.
J. Zhao et al. / Progress in Materials Science 83 (2016) 24–151 39

Usually, Z2 invariant is employed to classify the topological properties of insulators with TR symmetry, such as pristine
silicene. But in the present situation, the Chern numbers are used, because the Chern numbers (including spin-Chern num-
bers) of pristine silicene, have analytical simple expressions. The Chern numbers with spin (sz) and valley (g) indexes are:
g
C gsz ¼   Sgnðmgsz Þ; mgsz ¼ gtiso sz ; ð7Þ
2
where mgsz acts as the Dirac mass. In Eq. (7), one has not considered the intrinsic Rashba SOC (tiRso) since tiRso is relatively
small compared with tiso. Since the topological properties of system would not change without a band closing, which cannot
occur with small tiRso, the topological properties of pristine silicene with and without intrinsic Rashba SOC remain the same
[114]. With C gsz , we can introduce three nonequivalent Chern numbers [125–127]:
0 0
C ¼ C K" þ C K" þ C K# þ C K# ; ð8Þ
1 K 0 0

Cs ¼ C þ C K"  C K#  C K# ; ð9Þ
2 "
0 0
C t ¼ C K" þ C K#  C K"  C K# ; ð10Þ

where C is the total Chern number, and Cs(t) is the spin (valley)-Chern number. The Chern numbers of pristine silicene are
(C Cs Ct) = (0 1 0), corresponding to the QSHE. By analyzing the energy spectrum of a zigzag-terminated silicene nanoribbon,
it was found that on the boundaries of nanoribbon the helical edge states with spin up and down flow along the opposite
directions, respectively [114].

2.5.3. Extrinsic topological states


Except for the quantum spin Hall state, silicene also hosts various topological states under different external fields
[114–116]; thus, it holds great promises in nanoelectronic devices [31,32,42,128]. In the following, we will briefly illustrate
how these topological states emerge under external fields and discuss their topological properties by Chern numbers,
noticing that the TR symmetry of system is broken in some situations.
Electric field – Due to the low-buckled geometry, the external electric field is the most popular way to tune the electronic
properties of silicene [31,32]. The low-energy effective Hamiltonian around the valley points induced by electric field is:
Hgel ¼ Dsz ; ð11Þ
where D is the staggered sublattice potential induced by applying electric field perpendicular to the silicene sheet. In the
presence of electric field, the structure inversion symmetry of silicene is broken and the external Rashba SOC emerges.
But such induced Rashba SOC is rather weak and negligible [43]. Now the Dirac mass is mgsz ¼ gt iso sz þ D. When the strength
of electric field increases and reaches a critical value Dc ¼ tiso , spin- and valley-resolved Dirac cones appear (Fig. 15b). With
further increasing D, the band gap of silicene reopens at the Dirac points, as shown in Fig. 15c, and silicene undergoes a
topological phase transition, as the Chern numbers change from (C Cs Ct) = (0 1 0) to (0 0 2).
Exchange field and Rashba SOCs – Generally, silicene is grown on certain substrate. Due to the interaction with the
substrate, the electronic properties of silicene may show some differences from the pristine one [129,130]. Besides, the
influence of metal atom adsorption on silicene may be similar to that of substrates [33,131–134]. The exchange field can
be naturally introduced by ferromagnetic insulating substrate, or by adsorbing magnetic atom [114,135]. The low-energy
effective Hamiltonian of exchange field is:
Hgex ¼ Msz : ð12Þ

Fig. 15. Evolution of band structures of the bulk silicene under electric field (D) and exchange field (M). (a) D ¼ 0, M = 0; (b) D ¼ tiso , M = 0; (c) D ¼ 2tiso ,
M = 0; (d) D ¼ 2tiso , M = tiso.
40 J. Zhao et al. / Progress in Materials Science 83 (2016) 24–151

In the presence of both exchange field and electric field, spin and valley perfect polarized states can be obtained (Fig. 15d).
Another important feature induced by the proximity coupling to insulating substrate or the metal–atom adsorption is the
enhancements of the strengths of intrinsic and extrinsic Rashba SOC [33,131–134]. Thus, both exchange field and consider-
able Rashba SOC should be taken into account to describe the electronic properties of silicene. Although the Rashba SOC
would destroy the QSHE in silicene, they can establish another striking topological effect – the quantum anomalous Hall
effect (QAHE) [116].
The TB Hamiltonian and the corresponding low-energy effective Hamiltonian of extrinsic Rashba SOC read [114,116]:
2teRso X y
c ðr  dij Þab cjb ;
z
HeRso ¼ i ð13Þ
3 hi;jiab ia
HgeRso ¼ t eRso ðgsx sy  sy sx Þ: ð14Þ

With exchange field, the competition between intrinsic and extrinsic Rashba SOC may result in two distinct quantum
anomalous Hall states [116].
The evolution of the band structures of silicene and zigzag-terminated nanoribbon with the Rashba SOC and exchange
field is presented in Fig. 16. When the intrinsic Rashba SOC dominates the electronic properties, the Chern number of silicene
is C = 2, because there are two pairs of nontrivial gapless edge modes in the nanoribbon, localized in the two valleys respec-
tively. Since the two valleys have equal contribution to the Chern number, the corresponding valley Chern number
0
C t ¼ C K  C K is zero [116,136] (note that spin is not a good quantum number here). When the strength of the extrinsic
Rashba SOC becomes strong, the interplay between these two Rashba SOCs will result in closing and reopening of the gap
in one valley, while in the opposite valley it keeps increasing. Hence a band inversion appears around one valley (K0 ), with
0 0
Chern number of the valley changing from C K ¼ 1 to C K ¼ 2. Consequently, the valley Chern number is Ct = 3 now, indi-
cating the insulator in the present model can host both QAHE and quantum valley Hall effect (QVHE).
Electromagnetic field – The QSHE of pristine silicene can be also destroyed by a Haldane term [115,137,138], which may be
induced by applying off-resonant light [139–141]. Under the off-resonance coherent laser beam which does not directly
excite electrons, the influence of light can be described by a static effective Hamiltonian [115,139]:

A2 X
Hha ¼ iht2F pffiffiffi tij cyia cjb ; ð15Þ
3 3Xph hhi;jiiab

where the vector potential of circular polarized light is A(t) = (±A sin Xpht, A cos Xpht) with + () represents the left(right)
circulation and A ¼ eAa=  Xph  t indicates the off-resonant condition. The Haldane term
h, Xph is the frequency of light, and h
in the above model can be understood as a second-order process: electron emits (absorbs) a photon and then absorbs (emits)
a photon [139]. The expression of Hha in Eq. (15) can be also derived by Floquet theory, which is established to describe the
time-periodic systems [142,143]. The low-energy effective Hamiltonian of Haldane term is:

A2
Hgha ¼  h2 t2F gsz : ð16Þ
hXph

The Dirac mass of pristine silicene with electric field and Haldane term is mgsz ¼ gsz t iso þ D  hAX 
2 2
h t2F g, here we drop the
ph

Rashba SOC again. Due to the competition between the intrinsic SOC, electric field, and circular polarization light, the phases
 2

of silicene in the D; XA space are abundant [115].
ph
Ó American Physical Society 2016

Fig. 16. Evolution of the band structures of the bulk (a–f) and zigzag-terminated (g–l) silicene as a function of the extrinsic Rashba SOC at fixed intrinsic
Rashba SOC (tiRso = 0.08t) and exchange field (M = 0.5t). (a–f)/(g–i) tiRso/t = 0, 0.01, 0.031, 0.045, 0.067, 0.09, respectively. Colors are used to label the edge
modes localized at opposite boundaries. The vertical axis is the energy, and the horizontal axis is the momentum. Reproduced with permission from Ref.
[116].
J. Zhao et al. / Progress in Materials Science 83 (2016) 24–151 41

Magnetic field – When silicene sheet is exposed to a perpendicular magnetic field, the anomalous integer and half-integer
quantum Hall effect (IQHE and HIQHE) would emerge. It is known that the Hall conductivity in monolayer graphene shows
an exotic integer sequence 4(N + 1/2) in unit of conductance quanta, where the degeneracy 4 accounts for the spin degen-
eracy and valley degeneracy, and the half integer shift is caused by the p Berry phase for the massless Dirac fermion
[144,145]. Due to the strong SOC, the IQHE in silicene, compared with that in graphene, manifests some peculiar behaviors
[123,146–148]. For example, the spin intra-Landau level transitions, which are absent in graphene, could emerge in silicene
when a perpendicular electric field is applied [123,146,147]. These transitions lift the spin and valley degeneracy of the
Landau levels and lead to the HIQHE in silicene [123,146,148] (Fig. 17). Furthermore, if the external electric field is a period
field, a nearly perfect spin polarization for diffusive conductivity, which stems from the modulation broadened Landau
levels, can be obtained in certain ranges of magnetic field [147,148].

2.6. Carrier mobility and electron–phonon interaction

Carrier mobility is always a major concern in the research of electronic materials, as high carrier mobility leads to great
device performance in transport. The low-energy electrons of silicene are approximately considered as massless Dirac
fermions like those of graphene [1,144,149], despite there is a gap opened by SOC; thus, the carrier mobility of silicene is
expected to be high. But it is also very sensitive to the electron–phonon (e–ph) interaction. The strong e–ph interaction
can drastically reduce the carrier mobility of a system.
The carrier mobility is defined as ratio of the drift velocity td to the electric field E [150]:

t  jejs
 d m
lm  ¼ ; ð17Þ
E m

where e is the electron charge, sm is momentum relaxation time, and m⁄ is the effective mass of electron or hole.
The e–ph interaction affects the carrier mobility by reducing the momentum relaxation time sm. The e–ph interaction of
silicene is considerable, as the larger Si–Si interatomic distance of silicene, compared to graphene, weakens the p–p overlaps
and results in a low-buckled structure with sp3-like hybridization [10]. Thus, just like the hole-doped graphane [151], the
e–ph interaction of silicene has important influence on the transport behavior [76,152].
The intrinsic carrier mobility of silicene has been calculated by Shao and co-workers, where the deformation potential
theory is used to the electron–phonon interaction [92]. For silicene at room temperature, the obtained electron and hole
mobilities are 2.57  105 cm2 V1 s1 and 2.22  105 cm2 V1 s1, respectively, which are smaller than those of graphene
(3.20  105 cm2 V1 s1 and 3.51  105 cm2 V1 s1 for electron and hole, respectively) but still in the same order of magni-
tude. For comparison, the typical carrier motilities of bulk silicon are 1400 cm2 V1 s1 for electron and 450 cm2 V1 s1 for
hole, which are two to three orders of magnitude lower than that of silicene.
Later, the electron–phonon interaction in silicene was calculated from a more reliable first-principles method [76]. In
Fig. 18, the e–ph interaction matrix elements of six distinct phonon modes of silicene from first-principles calculations
are presented [76]. The e–ph interaction matrix elements of acoustic phonon modes, especially the ZA mode, are large. By
using full-band Monte Carlo simulation, the carrier mobility of silicene was estimated as approximate 1200 cm2 V1 s1
for electron at room temperature. This is two orders of magnitude smaller than that obtained by Shao et al. [92], reflecting
the serious deficiency of the deformation potential theory in treating silicene. If the ZA vibration are carefully dampened, the
electron mobility of silicene would be enhanced to 3900 cm2 V1 s1 [76]. For silicene nanoribbon, the carrier mobility
would be reduced but still large [153].
In the recently fabricated silicene field-effect transistor, the measured mobility for both electrons and holes is not high:
about 100 cm2 V1 s1 at room temperature [42]. This low mobility may be attributed to strong scattering from acoustic
Ó American Physical Society 2016

Fig. 17. Hall conductivity ryx and longitudinal component rxx as a function of electron density in silicene under magnetic field of B = 3 T and temperature of
T = 2 K: (a) Ez = 0, (b) elEz = 12 meV. Reproduced with permission from Ref. [123].
42 J. Zhao et al. / Progress in Materials Science 83 (2016) 24–151

Ó American Physical Society 2016


Fig. 18. Electron–phonon interaction matrix elements for all six modes. Reproduced with permission from Ref. [76].

phonons, especially the out-of-plane acoustic phonon [76], as well as possible existence of structural defects in silicene
sheet.

2.7. Superconductivity

Superconductivity (SC) has been one of the central issues of condensed matter physics and materials science for the past
century due to the rich physics and important applications [154–156]. Recently, the topological superconductor [157,158], a
counterpart of topological insulator, has attracted lots of attention, because its possible Majorana fermion edge modes can be
used as the building block for a topological quantum computer [159].
Generally, the topological SC may be induced by SC combined with topologically nontrivial state [160]. Accordingly,
silicene is a promising material for realization of quantum computation for several reasons. First, silicene has been proved
to have nontrivial topological properties. Second, the e–ph interaction of silicene is strong and can be further enhanced by
strain; thus, the conventional phonon-mediated SC is naturally expected to occur in silicene. Third, the electron correlation
effect of silicene may be strong and thus may also induce the unconventional SC, as the usual nearest-neighbor hopping of
silicene is not large (t = 1.08 eV) while the electron–electron interaction of silicene is less screened compared to the bulk 3D
materials. Moreover, a possible superconducting gap has been observed in silicene epitaxially grown on Ag(1 1 1) surface
[44]. In the following, we will discuss the superconducting properties of silicene, induced by electron–phonon and
electron–electron interaction, respectively.

2.7.1. Phonon-mediated superconductivity


Compared with the pristine one, silicene grown on the Ag(1 1 1) surface has two distinct features: shifted Fermi energy
induced by transferring electrons from Ag surface to silicene, and biaxial strain due to the lattice mismatch between Ag
lattice and silicene layer [44]. These two feathers both largely enhance the strength of e–ph interaction of silicene, as
discussed below. Thus, the electron-doped silicene was predicted to be a good 2D electron–phonon superconductor under
biaxial tensile strain by first-principles calculations [161].
In Fig. 19, the band structures of silicene near the Dirac point under biaxial tensile strains are presented [161]. When the
Fermi level is located at the position of Van Hove singularity, where silicene is still stable [162,163], the nesting effect would
be largely enhanced, indicating strong electron–phonon coupling. Fig. 19 also presents the projected destiny of states (PDOS)
for silicene under different tensions. In the position of Van Hove singularity, there exist components arising from the s, px and
J. Zhao et al. / Progress in Materials Science 83 (2016) 24–151 43

Ó IOP Publishing 2016


Fig. 19. Band structures of silicene near the Dirac point under the tension of (a–c) e = 0%, 3%, 5%. The Fermi level is set to zero. Projected density of states
(PDOS) for silicene under the tension of (d–f) e = 0%, 3%, 5%. The dashed line in band structure and PDOS is the location of Van Hove singularity. Reproduced
with permission from Ref. [161].

Ó IOP Publishing 2016

Fig. 20. Carrier concentration dependence of (a) k and (b) Tc of silicene under different tensions. Reproduced with permission from Ref.
[161].

py orbitals of silicon atoms in the PDOS induced by tension, which has very important influence on the e–ph coupling and SC.
Since pz orbital is antisymmetric with respect to the basal plane, the e–ph matrix element induced by pz orbital would be
zero if the sheet is planar [164]. Considering the decrease of buckling height for silicene under tension, the tension-
enhanced e–ph coupling in silicene is attributed to the increase of DOS from the s, px and py orbitals of silicon atoms at
the Fermi level.
In Fig. 20, the e–ph coupling constant k and superconducting critical temperature Tc as a function of carrier concentration
under different tension are presented. Figs. 19 and 20 intuitively show that more electronic states and weight of px and py
orbitals at the Fermi surface result in larger k and higher Tc. Accordingly, Tc of electron-doped silicene can be increased to
over 10 K by 5% tensile strain [161]. It is interesting that although the electron-doped silicene with biaxial tensile strain
can be a phonon-mediated superconductor, the thermodynamic properties of the superconducting phase were predicted
to differ from the predictions of the BCS theory [165].

2.7.2. Unconventional superconductivity


Usually, the unconventional superconductivity is related to cuprates, iron compounds, heavy-fermion superconductors,
and so on [154–156]. Silicene may also host various kinds of unconventional superconducting states, including time reversal
breaking chiral SC, time reversal invariant helical SC, and the general triplet SC [166–168]. In particular, the pairing symme-
try of SC in silicene can be tuned by electric field and doping [166,168]. In the following, we will discuss the chiral d + id0 SC
and equal-spin helical p + ip0 SC in undoped and doped bilayer silicene [167,168], as well as the electric-field induced quan-
tum phase transition to the f-wave superconducting state in doped monolayer silicene [166].
44 J. Zhao et al. / Progress in Materials Science 83 (2016) 24–151

Monolayer silicene – The doped silicene monolayer under a perpendicular external electric field mainly exhibits two kinds
of superconducting states with different pairing symmetry: singlet d + id0 chiral SC under weak field and the triplet f-wave SC
under strong field. Noticeably, Tc of the f-wave pairing state is controllable, and can be remarkably enhanced by applying
electric field [166].
The monolayer silicene in a perpendicular external electric field can be described by the following Hamiltonian [166]:
X X X
H ¼ t ðcyir cjr þ h:c:Þ þ D gi nir þ U ni" ni# ; ð18Þ
hi;jir ir i

where gi = 1/1 for A/B sublattice, and U represents the Hubbard interaction strength. The SOC terms are not considered
here. When D is large (D > 0), sublattice A will dominate over sublattice B, and be responsible for the main low-energy
physics. This sublattice symmetry breaking effect turns out to be crucial for the emergence of the triplet f-wave SC.
Within random phase approximation (RPA) for the present model [169–171], a critical interaction strength Uc can be
obtained. When the interaction strength U is greater than Uc, the spin susceptibility of the model diverges, which implies
an instability towards the long-range spin-density-wave (SDW) order. Below Uc, the spin fluctuations play the major role
in mediating the Cooper pairing. As the Hubbard term U is close to but still lower than Uc, the antiferromagnetic spin
fluctuation is strong. Through exchanging spin susceptibility, one obtains the effective interaction vertex (Vab(k, k0 )) between
two Cooper pairs on the Fermi surface (FS) [166], which is then plugged into the following linearized gap equation near Tc:
Z
1 X
0
0 V ab ðk; k Þ 0
 dkk b 0
Db ðk Þ ¼ kDa ðkÞ: ð19Þ
4p 2 b Fs t F ðk Þ

0 0
Here the integration is along the bth FS patch. tbF ðk Þ is the Fermi velocity and kk represents the component along the FS. By
diagonalizing this eigenvalue problem, the largest eigenvalue k and the corresponding eigenvector Da(k) can be obtained.
The eigenvalue k is related to the Tc (Tc  e1/k), and Da(k) determines the leading pairing symmetry of system.
The phase diagram on the plane of doping (x) and electric field (D) is shown in Fig. 21a [166]. Except for the SDW phase,
the singlet chiral dx2 y2 þ idxy and triplet nodeless f-wave pairings overcome the other instabilities and serve as the leading
instability in different regions of the phase diagram. Fig. 21b manifests that the gap function of the dxy symmetry (for doping
x = 0.15 and D = 0) on the FS is asymmetric about the x and y axes, while that of dx2 y2 is symmetric about these axes (not

Ó Macmillan Publishers Limited 2016

Fig. 21. Superconducting phase diagram, gap functions, critical interactions and pairing eigenvalues. (a) The superconducting phase diagram of the system
on the plane of doping (x) and electric field (D). Distributions of the gap functions on the FS for x = 0.15, (b) dxy symmetry for D = 0 and (c) f-wave symmetry
for D = 1 eV. The main parts of the real-space pairings for (d) d + id0 symmetry and (e) f-wave symmetry. (f) The doping dependence of the SDW critical
interactions Uc for D = 0 and D = 1 eV. (g) The electric-field dependences of the largest eigenvalues k of f-wave and d-wave pairings for x = 0.15. Inset: the
electric-field dependence of DOS at the Fermi level for x = 0.15. Reproduced with permission from Ref. [166].
J. Zhao et al. / Progress in Materials Science 83 (2016) 24–151 45

shown). The two d-wave pairings mix in the form of d + id0 minimizing the ground state energy [172]. In Fig. 21c, the gap
function of the f-wave pairing (for x = 0.15 and D = 1 eV) is presented. The main parts of the d + id0 and f-wave pairing in real
space are distributed on nearest-neighboring bonds and near nearest-neighboring bonds respectively (Fig. 21d and e).
The phase transition between d + id0 and f-wave pairing originates from the sublattice symmetry breaking effect, which
has selected the sublattice A as the low energy subspace of the system. All the relevant low-energy physics, the effective pair-
ing interaction mediated by these spin fluctuations, and the pairing itself take place mainly in this subspace under strong
electric field. Consequently, at low doping, the inter-sublattice pairing d + id0 symmetry shown in Fig. 21d has to give way
to the intra-sublattice pairing f-wave symmetry shown in Fig. 21e.
In particular, Tc of the f-wave pairing state can be significantly enhanced by external electric field, as shown in Fig. 21g.
Physically, such enhancement is attributed to the increase of DOS near the Fermi level (inset of Fig. 21g) due to flattening of
the band structure under the electric field. The increase of DOS not only increases the number of Cooper pairs near the FS, but
also enhances the pairing interaction. One more important feature of the f-wave pairing proposed here is that it needs
neither long-range Coulomb interaction nor the vanishingly small Hubbard U, and thus is more realizable than the other
proposals.
Bilayer silicene – Bilayer silicene (BLS) also exhibits two kinds of unconventional superconducting states, i.e., chiral d + id0
SC in the undoped and low doped region [167], and p + ip0 SC in the vicinity of the ferromagnetic regions around the Van
Hove doping levels [168].
First-principles calculations reveal two stable stacking ways of silicene layers, among which the energetically favored one
(named the AB-bt structure) is shown in Fig. 22a, and the corresponding phonon spectrum is shown in Fig. 22b. The band
structure of the AB-bt stacking BLS is shown in Fig. 23a, together with its low energy zooming. For undoped BLS, the most
obvious feature of the band structure is the 300 meV overlap between the valence band and the conduction band, much
larger than the 1.6 meV in the bilayer graphene and the 40 meV in the graphite [173,174]. The band crossings occur not only
at the K points, but also on the K–U axes with an energy difference between them. Such crossings result in a three-folded
symmetric pocket FS structure surrounding each K point, as displayed in Fig. 23b, where the central electron pocket is
accompanied by three identical outer hole pockets with equal total area.
It is important to note that the band structure of the BLS system is considerably sensitive to biaxial strains. As shown in
Fig. 24a, for small strains which keep the symmetry and FS topology of the system, the total area of the electron or hole pock-
ets has a considerable variation. This tunable property of the band structure turns out to be very important for the following
discussions.

Ó American Physical Society 2016

Fig. 22. (a) Optimized geometry of the bilayer silicene. (b) The corresponding phonon spectrum. In (a), both the top view (left) and side view (right) are
shown. The vertical bond length lt, the intralayer nearest neighbor bond length ln, and the angle h between them are marked, together with the hopping
integrals between each two of the four atoms A1(2) and B1(2) within a unit cell. Reproduced with permission from Ref. [167].
Ó American Physical Society 2016

Fig. 23. (a) The band structure of bilayer silicene corresponding to the optimized lattice structure in Fig. 22. (b) The FS patches around the K point. In (a), the
zooming in low-energy band (right) is also shown, where the TB model (red scatters) is compared with the first-principles results (blue solid line). In (b), the
central pocket (red) is electron-like and the outer three identical pockets (green) are hole-like, with the total areas of the two kinds of pockets equal.
Reproduced with permission from Ref. [167].
46 J. Zhao et al. / Progress in Materials Science 83 (2016) 24–151

Ó American Physical Society 2016


Fig. 24. (a) The biaxial strain dependence of Fermi pocket area ratio viz. the ratio of the total area of the electron and hole pockets against the total area of
the Brillouin zone and the SDW critical value Uc of the BLS. (b) The interaction strength U dependence of the largest eigenvalue k of the linearized gap
function, which is related to Tc through T c  e1=k . Results for different strain values are compared. (c) k dependence of the gap function of the dx2 y2
symmetry for U = 1 eV, which is asymmetric about the axes x = ±y shown in the reciprocal space. Inset: zooming in the vicinity of K0 . Reproduced with
permission from Ref. [167].

Similar to the case of monolayer silicene, the superconducting properties of bilayer silicene can be captured by the
following Hubbard model [167]:

X X
H¼ cykar Hab ðkÞckbr þ U nia" nia# ; ð20Þ
k;r;ab i;a

where H(k) is a four band low-energy effective Hamiltonian, which reflects all the low-energy features of BLS, i and a(b)
denote the unit cell and orbital indices, respectively.
The SDW critical value Uc can be obtained from RPA calculation. The leading pairing symmetries of the system for U < Uc
at all strain values are exactly the degenerate dxy and dx2 y2 doublets. Similarly, the two d-wave pairings mix in the form of
d + id0 in order to minimize the ground state energy.
The U dependence of the eigenvalue k, which is related to Tc through Tc  e1/k, is depicted in Fig. 24 for different strains.
Clearly, Tc increases with the Hubbard U term and rises promptly at U=U c K 1 as a result of the strongly enhanced antifer-
romagnetic spin fluctuation near the critical point. Since Uc is tunable via strain, as shown in Fig. 24, the U/Uc ratio varies
within a range that provides the basis for the condition of the relation U=U c K 1, which is crucial for the high Tc of the system.
For example, for k  0.3 attainable by different strains shown in Fig. 24b, the obtained Tc can be as high as 80 K, although it is
usually overestimated in the RPA level. For real material, whether high Tc can be acquired is determined by how near U/Uc
can be tuned to 1.
For doped BLS, the leading instability of the system is itinerant ferromagnetism in the narrow doping regions around the
Van Hove singularities, and triplet p + ip0 superconductivity with a possible high Tc emerges in the vicinity of these regions
[168], as displayed in Fig. 25. When an extra weak Kane–Mele SOC is added to the system, the equal-spin helical p + ip0
pairing will dominate over the chiral one and serve as the leading instability. This intriguing triplet superconducting state
has TR invariant weak topological property, and can harbor the Majorana zero mode at its boundary.

Fig. 25. The doping dependences of the pairing eigenvalues k for all possible pairing symmetries. The vertical bold gray lines indicate the doping regions
where the itinerant FM occurs. Inset: the typical split between the helical ðpx þ ipy Þ"" , ðpx  ipy Þ## pairing and the chiral ðpx  ipy Þ"##" pairing caused by the
weak SOC term. The split in other doping regions where the p + ip0 SC occurs is similar to the one shown in the inset [168].
J. Zhao et al. / Progress in Materials Science 83 (2016) 24–151 47

Ó American Physical Society 2016


Fig. 26. (a) Filled-states scanning tunneling microscopy (STM) image of the 2D Si layer on Ag(1 1 1)-(1  1) (Ubias = 1.3 V, I = 0.35 nA). Clearly visible is the
honeycomb-like structure. (b) Line profile along the dashed white line indicated in (a). The dark centers in the STM micrograph are separated by 1.14 nm,
corresponding to four times of the Ag(1 1 1) lattice constant, in agreement with the (4  4) symmetry. (c) High-resolution STM topography (3  3 nm,
Ubias = 1.3 V, I = 0.35 nA) of the Si adlayer. (d) Fully relaxed atomic geometries of silicene on the Ag(1 1 1) surface with (4  4) superstructure from DFT
calculation. (e) Simulated STM image for the structure shown in (d), which exhibits the same structural features as those observed in the experimental STM
image in (c), i.e., a hexagonal arrangement of the triangular structure around dark centers. Reproduced with permission from Ref. [19].

3. Synthesis and characterization of silicene

Despite the theoretical prediction of silicene down to 2004, the breakthrough in experimental synthesis of silicene took
place in year 2012, when several groups, including the Le Lay group [19], Takagi and Kawai group [21], Wu group [20,175],
Aufray group [22], Oughaddou group [176], reported in parallel the successful preparation of monolayer silicene sheets on Ag
(1 1 1). It should be noted that the successful synthesis of silicene on another substrate, ZrB2(0 0 0 1), was achieved indepen-
dently by Fleurence et al. around the same time [27]. One should also be aware that Laimi et al. had claimed the successful
growth of silicene on Ag(1 1 1) in an earlier paper at 2010 [177]. But their results were not reproduced by the other groups
and the reported Si–Si distance (0.19 ± 0.01 nm) was unreasonably too short. Following the successful preparation of
silicene, the studies on silicene sees a sudden burst in the following few years.
Silicene on various substrates usually display complicated surface superstructures, in contrast to graphene which usually
displays honeycomb, none-reconstructed surface. The mechanism for the formation of these superstructures is due to the
buckling of silicene lattice, which will rearrange due to the interaction with the substrate. As a representative, the STM image
of silicene sheet on Ag(1 1 1) surface with (4  4) superstructure reported by Vogt et al. [19] is shown in Fig. 26. In this struc-
ture, silicene sheet sits on top of Ag(1 1 1) surface in such a way that a (3  3) silicene supercell coincides with a (4  4)
supercell of the Ag(1 1 1) substrate, forming a so-called (4  4) superstructure. Fig. 26b and c shows that the dark centers
in the STM micrograph are separated by an average distance of 1.14 nm, corresponding to four times of the surface Ag
(1 1 1) lattice constant. As shown in Fig. 26d, each rhombus consists of six bright spots in the STM image, corresponding
to six protruding Si atoms in each (3  3) silicene unit cells that can be explained by the buckled structure of silicene on
Ag(1 1 1) surface.
In this section, we will discuss in detail the current status of experimental synthesis and characterization of silicene on
different substrates, from Ag(1 1 1), ZrB2(0 0 0 1) to others. We will cover the growth mechanism, surface structures and
electronic properties, which are currently the focusing issues in the experimental studies of silicene. The relevant theoretical
48 J. Zhao et al. / Progress in Materials Science 83 (2016) 24–151

results from first-principles calculations on the interaction between silicene and substrates will also be briefly included for
convenience of discussion.

3.1. Silicene monolayer on Ag(1 1 1) substrate

The fabrication of silicene film has to be conducted in ultrahigh vacuum (UHV) environment due to the sensitivity of Si
upon oxidation. Owing to the clean environment and easy control of the deposition coverage with sub-monolayer precision,
molecular beam epitaxy (MBE) is the preferable approach to fabricate silicene sheets via epitaxial growth of silicon on solid
substrates. A clean Ag(1 1 1) substrate can be achieved by repeated Argon ion sputtering and annealing process. Silicon is
deposited on substrate by thermal evaporation from a pure Si source. The simplest Si source involves a silicon bar cut from
a silicon wafer and clamped between two electrodes. As silicene has sufficiently high vapor pressure below the melting tem-
perature, stable silicon flux can be obtained when the silicon bar is heated up passing a direct current it. In contrast, germa-
nium source cannot be constructed in this way because its vapor pressure is very low below the melting temperature.
Instead of single crystal Ag(1 1 1), recently, Ag(1 1 1) films prepared on mica substrate were also used as the substrate for
silicene growth [42]. The advantage of using Ag films instead of single crystalline Ag(1 1 1) is that the Ag films can be easily
peeled off from the mica substrate and etched away, enabling the fabrication of a silicene device. Moreover, Ag(1 1 1) film is
on mica is commercially available with large size, at much lower price than that of single crystal Ag(1 1 1), making it a
promising substrate for mass production of potential silicene devices.
Depending on the Si coverage and substrate temperature, silicene can form a variety of superstructures on Ag(1 1 1)
surface [19–23,175,176,178–183]. At lower substrate temperature below 400 K, silicon atoms deposited on Ag(1 1 1) tend
to form clusters or disordered structures [20]. When the substrate temperature is above 400 K during deposition, the silicene
p p p p p p
sheet exhibits typically five ordered phases: from T phase to 4  4, ( 13  13)R ± 13.9° ( 13  13 for short), 2 3  2 3,
p p
and finally to a 3  3 phases with increasing substrate temperature. In addition to the above mentioned phases, other
ordered structures of silicene were occasionally reported [177,184–186]. The fundamental reason for silicene to exhibit
so many reconstructions, as compared with the none-reconstructed graphene structure, is due to the buckling mechanism
in silicene. Although the honeycomb lattices of silicene are all preserved on Ag substrate in these phases, the buckling pat-
terns are rearranged in these phases, and thus they show apparently distinct STM images [24]. We will discuss the details of
the structures and buckling patterns of each phase in the following sections.
To date, the reports of different phases of the silicene grown on Ag(1 1 1) and discussions of their structural models in
p p
literature remain controversial. For instance, the T phase was sometimes ascribed to a special 13  13 phase [187],
p p p p
the structure of the 13  13 was under debated, and the 2 3  2 3 phase was described as ‘‘highly defective and
disorder structure” [184]. Even more, whether alloying occurs during growth [187–189] and whether the Dirac state exists
are still under debates [190–199]. In the following, we intend to give a comprehensive picture on the growth and structures
of various silicene structures on Ag(1 1 1), covering all above debating issues. The structure models based on STM observa-
tions and first-principles calculations will be systematically described.
In literature, the notation of these reconstruction phases is referred to both the pristine silicene-1  1 lattice and the Ag
p p p p
(1 1 1)-1  1 lattice. Therefore, the 4  4 phase is also cited as 3  3, and the 2 3  2 3 phase as 7  7 with respective to
p p p p
the silicene-1  1. To avoid possible confusion, we will occasionally use Si(3  3)/Ag(4  4) and Si( 7  7)/Ag(2 3  2 3)
in the following discussions. The term of silicene 1  1 as reference is sometimes useful if the film is not commensurate with
the substrate.

3.1.1. T phase
The T phase was reported first by Feng et al. [20] and reproduced by other experiments [23,200,201], but rarely discussed
in literature. In STM images, it appears as big and round protrusions arranged in a hexagonal closed pack pattern. The
distance between neighboring protrusions is about 1.0 nm, as shown in Fig. 27. Although the exact atomic structure of this
phase is not clear yet, the following experimental facts are helpful to clarify its structural model.

(1) This phase does not show up in the same temperature regime with the other ordered structures. It emerges at the low-
est temperature regime where ordered Si phases just appear on Ag surface. Although this phase frequently coexists
p p
with 4  4 and 13  13 phases, it can be removed by annealing at appropriate temperature, leaving only 4  4
p p
and 13  13 phases on the Ag(1 1 1) surface. Since the thermal stability of this phase is lower than the 4  4
p p p p
and 13  13 phases, it shall not be included into the 13  13 phase.
(2) The periodicity of this phase is different from the other ordered phases. Apart from the hexagonal close packing of big
protrusions reported in literature, other periodic arrangement of the big protrusions, e.g., the parallel linear chains
shown in Fig. 27b, was also observed. This parallel chain structure should not be considered as a different phase since
it has the same formation condition and the same ball-like ‘‘basic unit” as the hexagonal one.
(3) Although normally this phase is imaged as big round protrusions in STM images, one can frequently see disorder in the
protrusions in the atomic resolution images, which can be recognized as two or three smaller protrusions with differ-
ent rotations [20]. This supports that this phase is likely a partially disordered phase consisting of locally incomplete
p p
silicene fractions and can be regarded as a ‘‘precursor phase” for more ordered phases like 4  4 and 13  13.
J. Zhao et al. / Progress in Materials Science 83 (2016) 24–151 49

Ó IOP Publishing 2016


Fig. 27. The T phase formed at relatively low substrate temperature. (a) (12  20 nm2) a surface with mostly hexagonal close packed structure of this phase,
p p
where small pieces of 13  13 phase are nucleated inside the T phase. (b) (20  50 nm2) a surface with better ordered T phase, in some area appearing as
long chains. (c) Top and side views of a silicon mono-layer structure with a coverage factor of 0.75 on a Ag(1 1 1) surface. Si atoms are shown with gray (dark
yellow) and dark gray (red) spheres. The panel (c) was reproduced with permission from Ref. [202].

Ó American Physical Society 2016

Fig. 28. The high resolution STM image and structural model of clean Si(3  3)/Ag(4  4)-a phase. (a) A typical STM image (14  11 nm2). In the upper-right
part, there is a small area consisting of a metastable 4  4-b phase. The white rhombus and the red rhombus mark the 4  4-a and 4  4-b unit cells,
respectively. (b) Structural model of 4  4-a. Each unit cell consists of six upper-buckled Si atoms and the two HUCs are mirror-symmetric. (c) Structural
model of 4  4-b. The orange rhombus is the unit cell drawn in Ref. [203], while the green rhombus is the redrawn unit cell. Reproduced with permission
from Ref. [203].

In literature, several structural models have been proposed for this phase. Wu’s group [20] ascribed this T phase as the
‘‘precursor” for the formation of 4  4 phase. The structure consists of incomplete pieces of silicene flakes rather than a com-
plete honeycomb lattice. In Ref. [20], they proposed a structural model where each protrusion in the STM image corresponds
to a hexagonal silicon ring. However, such separated silicon ring structure without hydrogen termination is not stable, as
proved by the binding energy calculation by Kaltsas et al. [202]. On the other hand, another incomplete silicene lattice model
that has higher silicon coverage (0.75 ML) was found as a local energy minimum configuration, even in the absence of any
hydrogen passivation of the Si atoms. This structure likely corresponds to the hexagonal packing structure of the T phase
(but not the linear chain structure). Interestingly, note that in this structure the hexagonal Si rings are deformed, and the
buckling of Si atoms are partially relaxed to form a more flat structure compared with freestanding silicene.
There are also theoretical models where this phase is described as a complete silicene-1  1 lattice overlapping on Ag
(1 1 1). For examples, Chiappe et al. [23] ascribed this phase as a complete silicene lattice overlapping on Ag(1 1 1) substrate
p p p p
forming either 13  13 or 7  7 periodicity. However, they did not explain the origination of the protruding features
in the STM images, but rather made correspondence between the STM image and the Moiré patterns formed by overlapping
50 J. Zhao et al. / Progress in Materials Science 83 (2016) 24–151

p p
Fig. 29. (a) (30  30 nm2) typical STM image of a surface with mix phases of 4  4 and 13  13. The cross lines indicates the relative rotation of the
orientation of the two phases. (b) (50  50 nm2) in the same image three domains are marked by numbers 1–3, where different long-range domain
boundary patterns can be recognized. Reproduced with permission from Ref. [178]. Copyright 2000–2016, American Scientific Publishers.

the silicene and Ag(1 1 1) lattices. Arafune et al. described it as 3.5  3.5R26° reconstruction, but without detailed analysis
p p
[201]. Liu et al. suggested that this phase has a periodicity and orientation identical to 13  13 phase, and thus ascribed
p p
it as a special 13  13 phase [200]. However, due to the experimental observation of complicated appearances of this
phase as well as the disorder in STM images, it is not suitable to describe this phase by a single model. We propose that this
phase is partially ordered, and may feature simultaneously rearranged buckling and incomplete silicene lattice. Therefore, it
is suitable to refer it as a ‘‘precursor state” that form in a relatively low temperature range.

3.1.2. Si(3  3)/Ag(4  4) phase


At relatively low substrate temperature, the first perfectly ordered phase observed in experiments is 3  3 phase with
respect to silicene 1  1, or 4  4 phase with respect to Ag(1 1 1)-1  1 substrate. Among all the monolayer silicene phases
on Ag(1 1 1), the 4  4 phase can be most easily recognized in STM images; thus it was independently reported by different
J. Zhao et al. / Progress in Materials Science 83 (2016) 24–151 51

Ó IOP Publishing 2016


p p
Fig. 30. (a) High-resolution topography STM image taken on a 13  13 domain showing a large hexagonal pattern. U = 3 mV and I = 0.2 nA. (b) Enlarged
STM image cut from the square indicated in (a). (c) Ball structural model for the vortex pattern shown in (b). Reproduced with permission from Ref.
[200].

Ó IOP Publishing 2016

p p
Fig. 31. Summary of the atomic structural models proposed previously for the 13  13 silicene/Ag(1 1 1) superstructure. Below each ball model is its
p p
corresponding STM simulation result with partial superimposition. (h) Schematic illustration showing two sets of 13  13 unit cell with rotational
 0 direction. Reproduced with permission from Ref. [200].
angles of ±13.9° from the ½1 1

groups in those early experiments [19–23,176]. As displayed in Fig. 28, the STM image of 4  4 phase exhibits triangular half
unit cells (HUC) each consisting of three bright spots, which is quite similar to the well-known Si(1 1 1)-5  5 surface
structure.
The 4  4 structure can be easily understood by the so called ‘‘magic mismatch” between the lattice constant of Ag
(aAg = 2.88 Å) and Si (aSi = 3.84 Å), i.e., 4  aAg = 3  aSi within 0.5% of error. Since both Ag(1 1 1) surface and silicene are
hexagonal lattices, a commensurate Si(3  3)/Ag(4  4) superstructure can be constructed by simply overlapping a
low-buckled, freestanding silicene 1  1 lattice on top of a Ag(1 1 1)-1  1 lattice in the same orientation. Such a magic
commensuration results in minimal interface strain and thus stabilizes the silicene structure.
52 J. Zhao et al. / Progress in Materials Science 83 (2016) 24–151

In the case of graphene on metal surfaces [204,205], overlapping two periodic lattices results in the famous ‘‘Moiré
patterns”. But in the Si(3  3)/Ag(4  4) phase, STM image shows six protrusions in each unit, not corresponding to a Moiré
pattern. Each protrusion originates from a real Si atom with higher position. In the freestanding silicene, there are two sets of
sublattices with different heights due to buckling. In each Si(3  3)/Ag(4  4) unit cell, there are 18 Si atoms. Originally, half
of them are upper buckled and half are lower buckled. Once the silicene-1  1 sheet is placed on a Ag(1 1 1) substrate, the
buckling configuration will spontaneously rearrange to lower the surface energy. As a consequence, there will be six
upper-buckled Si atoms per unit cell, while the other twelve Si atoms are all lower-buckled, as depicted in Fig. 28b.

Ó American Chemical Society 2016

p p
Fig. 32. (a) A derivative STM image (200  200 nm2, Vtip = 1.43 V) of a surface fully covered by the 2 3  2 3 phase. (b) High-resolution STM image
(15  15 nm2, Vtip = 1.0 V) showing the atomic structure. The bright areas exhibit complete honeycomb rings with a period of 1.0 nm, while other areas are
p p
disordered. (c) dI/dV spectra shows a peak at 0.3 V and a shoulder at 0.9 V are observed. (d) Calculated model of 2 3  2 3 superstructure of silicene. The
gray, yellow, and red balls represent the silver, lower silicon, and higher silicon atoms, respectively. (e and f) Experimental and simulated STM images
(1.0 eV above Fermi energy) showing the similar structure features and unit cell of lattice. Reproduced with permission from Ref. [20]
J. Zhao et al. / Progress in Materials Science 83 (2016) 24–151 53

Moreover, the three upper-buckled Si atoms in one HUC do not belong to the same sublattice with the three upper-buckled Si
atoms in the other HUC. This is in contrast to the case of freestanding silicene where all upper-buckled (or lower-buckled) Si
atoms constitute a sublattice. That the protrusions observed in STM correspond to upper-buckled Si atoms that are indeed
geometrically higher can be further confirmed by atomic force microscopy (AFM) measurements [180].
In addition to the normal 4  4 phase (denoted as 4  4-a), one can often observe another phase with a 4  4 periodicity,
namely the 4  4-b phase [200]. It emerges in the domain boundaries between a-4  4 phases probably because that this
phase is less stable than the normal 4  4-a phase and that the strain in domain boundary helps to stabilize it. The STM
images in Fig. 28a reveal that the two HUCs of the 4  4-b phase are different: one with six spots, another with only one spot
in the center.
The 4  4-b silicene structure was first reported by Liu et al. [200]. It was found that the STM image of the observed
4  4-b silicene could be explained nicely by a theoretical work [203], as shown in Fig. 28c. Note that in Ref. [203] the unit
cell was drawn in the orange rhombus and it was difficult to see the relation between 4  4-a and 4  4-b phases. If we
redraw the unit cell in the green rhombus, one can see that the two half-unit cells are inequivalent, one with six bright spots
and the other with one bright spot. The two 4  4 phases correspond to different buckling of the Si atoms, and all the bright
spots in the STM images correspond to upper buckled Si atoms in the silicene lattice.

p p
3.1.3. 13  13R13.9° phase
p p p p
The 13  13R13.9° ( 13  13 for short) phase usually coexists with the 4  4 phase [20,21,23,180,182,183,200,2
06–208]. With appropriate Si coverage, such a mix-phase monolayer film can fully cover the substrate surface. Although
the area ratio of these two phases can vary with the annealing temperature, it is usually unlikely to obtain a single-phase
p p
surface. This fact suggests that 4  4 phase and 13  13 have comparable thermal stability. Also, during the annealing
process, the total area of the silicene film does not show a notable change, meaning that the densities of Si atom in the
p p
4  4 and 13  13 are about the same. This is consistent with the model that both structures correspond to a complete
silicene lattice, with the different superstructures originating only from the rearranged buckling of the Si atoms.

p p
Fig. 33. (a) High resolution STM image of the (2 3  2 3)R30° structure with periodic perfect areas surrounded by disordered areas. (b) The structure
p p
model proposed by Jamgotchian et al., with a big unit cell (blue diamond) having a ( 133  133)R4.3° superstructure [212].
54 J. Zhao et al. / Progress in Materials Science 83 (2016) 24–151

p p p p
The typical STM images of 13  13 phase are presented in Fig. 29, showing the coexistence of 4  4 and 13  13
p p
phases. When the STM image resolution is not high enough, the two phases looks quite similar, and the 13  13 phase
p p
also has a dark corner hole and triangular unit cells. However, as marked in Fig. 29a, the orientation of 13  13 phase is
p p
rotated 13.9° with regard to the 4  4 phase. Note that in sufficiently large 13  13 domains there usually exist long-
range periodic patterns, which are similar to the Moiré patterns at a first glance. Many different long-range patterns have
been observed and a few examples are given in Fig. 29b.
p p
High resolution STM images [200] reveal the structure of 13  13 phase as well as its ‘‘Moiré-like” pattern, see
p p
Fig. 30a. Within a 13  13 unit cell (marked by rhombus), the most prominent feature is a trimer in the center of one
HUC, and a protrusion at the corner. The large ‘‘Moiré-like pattern” with periodic vortex structure is indeed a domain wall
structure, with each vortex surrounded by six triangular domains. The unit cell directions in neighboring domains are
rotated by an angle of 60°. It should be emphasized again that the whole surface is a complete and continuous honeycomb
silicene lattice. The appearance of these vortex patterns is solely due to the rearrangement of the buckling configuration of
silicon atoms.
p p
Till now, there have been several theoretical models proposed for the 13  13 phase [21,22,24,176,190,197]. As sum-
marized in Fig. 31, both model (a) and (e) produce STM images quite similar to the experimental ones, while the difference
between them is simply an additional buckled Si atom (red ball). Considering that the upper buckled Si atoms and the lower
buckled Si atoms should naturally take alternative positions in a silicene lattice, model (a) is more reasonable. In model (c),
all the Si atoms in a HUC are lower-buckled, and thus this model should be energetically less favorable.

p p p p
3.1.4. Si( 7  7)/Ag(2 3  2 3) phase
p p p p p p
In contrast to the coexistence of 4  4 and 13  13 phases, the Si( 7  7)/Ag(2 3  2 3) phase (denoted as
p p
2 3  2 3 thereafter), which was reported first by Feng et al. [20] and subsequently observed by many other groups
[22,23,176,179,181,183,209–211], is the only superstructure that can be prepared with single phase spreading over the
p p p p
entire Ag(1 1 1) surface. By slightly annealing a 4  4 and 13  13 mix-phase surface, a complete 2 3  2 3 phase
can be derived without apparent change of the area of the film. This indicates that the densities of Si atoms in the 4  4,
p p p p p p
13  13 and 2 3  2 3 structures are comparable, while the 2 3  2 3 phase has a higher thermal stability.

Ó American Physical Society 2016

Fig. 34. The STM image of a large area (65  65 nm2) consisting of a sheet of silicene on Ag(1 1 1) crossing two substrate steps. (b) The line profile as
indicated by the black line in (a) shows that the island is of one atom thick. (c) The high-resolution STM image (10  10 nm2) of the silicene surface taken at
tip bias 1.0 V. The honeycomb structure is clearly observed. (d) The line profile as indicated by the black line in (c) showing both the lateral and vertical
corrugation of the structure observed by STM. Reproduced with permission from Ref. [175].
J. Zhao et al. / Progress in Materials Science 83 (2016) 24–151 55

Ó American Physical Society 2016


p p
Fig. 35. (a and b) The STM images of the same area on ( 3  3)R30° silicene taken at tip bias 2.0 V and 0.1 V, respectively, at 5 K. (c) The filtered high-
p
resolution STM image with high contrast taken at 0.1 V. (d and e) Models of two energy-degenerated 3 reconstructed structures of silicene sheet on Ag
(1 1 1) surface with an orientation angle of h = 30°, which are obtained from DFT. Color code: Blue, yellow, and red spheres denote Ag atoms, Si atoms in
p
lower layer, and Si atoms in higher layer, respectively. The red triangles denote the units of 3 silicene structures. (f) The interpolated potential energy
p
curve for structural transition between the two the mirror-symmetric 3 geometries on Ag(1 1 1). (g) The intermediate structure between the two rhombic
p
3 structures of silicene shown in (d and e). Reproduced with permission from Ref. [218].

Recently, Liu et al. claimed that this phase is too defective to be thermodynamically stable [184]. Here we emphasize that
p p
2 3  2 3 phase is thermodynamically stable. However, it is challenging to elucidate its atomic structure, especially in the
apparently disordered area. Recently, Tao et al. reported the FET devices based on monolayer silicene prepared on and then
p p
peeled off from Ag(1 1 1) substrate [42]. In their experiment, silicene sheets with 4  4 and (2 3  2 3)R30° phases were
selected as the starting materials and the resulting devices exhibit basically the same transport behavior. If 4  4 silicene
p p
is an ordered phase while (2 3  2 3)R30° is severely defective, their transport properties such as carrier density and
mobility should be rather different.
p p p p
Fig. 32 shows typical large-area and zoom-in STM images of the Si( 7  7)/Ag(2 3  2 3) structure. Although the
prepared atomic-thick silicon film is entirely uniform in a large scale, the atomically resolved structure appears rather
defective and disordered. The large-size image displays periodic pattern of brighter areas, while the zoom-in image shows
that the brighter region consists of a few perfectly arranged hexagonal rings. On the other hand, the areas in between the
brighter area are rather disordered. With the aid of first-principles calculations, the atomic structure model of the perfect
p p p p
area can be established [20], which involves the Si- 7  7 super cell matching to the Ag-2 3  2 3 super cell. Within
p p
each Si- 7  7 super cell, there is one Si atom sitting exactly on top of a Ag atom that corresponds to the brighter spot.
p
These spots form a honeycomb lattice with period of 0.38  7 = 1.0 nm, in line with experimental observation. The 30°
p p  0i direction of Ag(1 1 1) is also confirmed by
angle between the lattice direction of the 7  7 superstructure and h1 1
experiments.
Careful examination of the disordered areas between the ordered areas reveals a hidden order (Fig. 33a). Around a hole
site there are always six spots, and connecting these six spots would result in a hexagon. Some of these six spots appear
darker, that is why the disordered area appears darker than the perfect area. Therefore, the large period pattern can be prob-
ably attributed to different electronic states in the ordered and disordered areas rather than the Moiré pattern. Recently,
p p
Jamgotchian et al. [212] proposed a structural model for 2 3  2 3 phase. As presented in Fig. 33, surrounding the perfect
area are distorted hexagonal rings that correspond to disordered area. Again, we emphasize that the seemingly disordered
area is not completely disordered. Most likely, the low-buckled honeycomb structure of silicene is distorted to accommodate
the strain, but the basic honeycomb network is still preserved.

p p
3.1.5. ( 3  3)R30° phase of silicene
p p
As substrate temperature reaches 500 K, silicene islands with ( 3  3)R30° periodicity (with respect to Si-1  1) are
observed on the surface. This phase is of particular interest because of the following two reasons: (i) it exhibits a metallic
56 J. Zhao et al. / Progress in Materials Science 83 (2016) 24–151

Ó Macmillan Publishers Limited 2016


Fig. 36. (a) Atomic structures, (b) local density of states (LDOS), (c) on-site Mulliken charges (|e|), (d) charge differential densities of Si24@Ag(1 1 1) system.
The yellow balls and sky blue balls represent Si and Ag atoms, respectively. The LDOS comes from the central six silicon atoms (3p orbital) highlighted by
red balls as well as the Ag atoms (4d orbital) highlighted by green ball right underneath the Si hexagon. In (d), red (blue) zone loses (gains) charge.
Reproduced with permission from Ref. [24].

surface state which generates quasi-particle interference (QPI) patterns (see Section 3.1.7); (ii) the surface of the multilayer
p p
silicene film exhibits exactly the same ( 3  3)R30° structure [20,179,180,213–217], which will be further discussed in
Section 3.4.
The STM image in Fig. 34 shows a silicene island across the step edges of the underlying Ag(1 1 1) surface without losing
continuity. The high-resolution image in Fig. 34c displays the honeycomb structure of silicene terrace. The lattice period of
p p
the honeycomb structure is about 0.64 nm, corresponding to a ( 3  3)R30° honeycomb superstructure with respect to
the 1  1 lattice.
p p
To explain the atomic structure of ( 3  3)R30° phase, we should mention the characteristic phase transition at low
temperature. When the silicene sample on substrate is cooled to liquid helium temperature (5 K), a dramatic phase transi-
tion occurs. As depicted in Fig. 35, one can see the emergence of atomic chains forming interconnected triangles, which are
actually boundaries separating two symmetric domains. At 77 K, the two neighboring protrusions in each honeycomb unit
cell are equally bright. After the phase transition (at about 40 K), one of them becomes much brighter than the other. Since
there are two possible configurations, the silicene film is phase separated into triangular domains with the two symmetric
configurations divided by narrow domain boundaries.
p p p p
The ( 3  3)R30° superstructure can be then explained by double symmetric ( 3  3)R30° reconstruction structures
p p
with dynamic flip-flop motion at high temperature, as shown in Fig. 35d–g. If ignoring the substrate, these two ( 3  3)
R30° superstructures have identical geometry and share the same central substrate atom for each hexagon unit. In every
p p
( 3  3)R30° unit cell, only one Si atom is buckled upward, whereas the rest five Si atoms have nearly identical lower
p p p p
height, resulting in a rhombic ( 3  3)R30° superstructure. The two types of rhombic ( 3  3)R30° superstructures
coincide well with the two mirror-symmetric phases observed in the low temperature experiments. At higher temperature,
p p
the flip-flop motion between the two phases leads to an averaged, symmetric ( 3  3)R30° phase.
p p
However, there is debate on the existence of ( 3  3)R30° structure in monolayer silicene. Is this silicene film mono-
p p
layer or multilayer (at least bilayer)? In an early paper, Wu et al. claimed that this ( 3  3)R30° structure exists in both
p p
monolayer and multilayer silicene films [20]. In their experiments, only the ( 3  3)R30° phase was observed after growth
of silicene film on Ag(1 1 1) at relatively high temperature. Successive experiments by several other groups [25,26,201]
p p
obtained the same ( 3  3)R30° phase by continue growing Si on top of a mix-phase (4  4) surface at a relatively low
p p
temperature, and islands of ( 3  3)R30° surface will form on top of the 4  4 film. These experimental facts indicate that
p p p p
2 ML and multilayer silicene films all exhibit a 3  3 surface structure, and that the previously reported ( 3  3)R30°
p p
phase was likely be a 2 ML film. However, one cannot explicitly exclude the possibility that monolayer ( 3  3)R30° may
p p
still exist, but rarely be observed, since the monolayer ( 3  3)R30° structure is quite stable from theoretical calculations
J. Zhao et al. / Progress in Materials Science 83 (2016) 24–151 57

Fig. 37. Isosurface plots of electron density distribution of the critical states which has contribution from the states of silicene’s p band (a–d) and p⁄ band
(e–h) on the K point. The isovalue is 0.02 e/Å3. Reproduced with permission from Ref. [195]. Copyright 2015, Elsevier B.V.

and was actually observed on other substrates like Ir(1 1 1), ZrB2 and ZrC [27–29]. In a theoretical study by Cahangirov et al.,
p p
it was suggested that the ( 3  3)R30° structure is constituted by dumbbell units of Si atoms arranged in a honeycomb
pattern [219].

3.1.6. Interaction with Ag(1 1 1) and growth mechanism


As discussed above, Ag(1 1 1) is the most frequently used surface for the epitaxial growth of silicene at present. Thus,
elucidating the interaction mechanism between silicene and Ag substrate is of significant importance for fundamental
understanding of the experimentally prepared silicene as well as the future development of silicene devices. To date, there
have been numerous theoretical calculations on the interactions between silicene sheets and Ag(1 1 1) surface [24,190,19
1,194–196,199,202,219–231], focusing on the hybridization between Si and Ag, the atomic structures of reconstructed
silicene phases, the electronic band structure and existence of Dirac cone, and the structural evolution of silicene film versus
coverage.
To illustrate the Si–Ag interaction and its implications on the growth mechanism of silicene, Gao and Zhao considered a
series of planar SiN clusters (i.e., silicene patches) on Ag(1 1 1) surface [24]. As shown in Fig. 36, a 2D Si24 cluster composed of
seven hexagonal rings (which can be viewed as a small patch of silicene) is placed on Ag(1 1 1) substrate and remains stable
upon relaxation. Strong hybridization between Si pz electrons and Ag 4d electrons is revealed by the LDOS for Si24@Ag(1 1 1)
(Fig. 36b), which shows that the 3p states of Si atoms couple with the 4d states of Ag atom in the vicinity of the Fermi level.
The strong p–d hybridization of Si–Ag can stabilize silicene sheet on Ag(1 1 1) surface, which is further supported by recent
Raman observation in experiment [42]. Such strong hybridization may account for the vanishing of Dirac fermions charac-
teristics of silicene on Ag(1 1 1) surface and will be discussed later. The charge differential densities of Si24@Ag(1 1 1) in
Fig. 36d illustrate that all Si atoms at both periphery and inner regions of Si24 cluster interact evenly with Ag(1 1 1) surface.
The homogeneous interaction between silicon atoms and Ag substrate is further supported by the homogeneous on-site
charges of Si atoms in Si24@Ag(1 1 1) system (Fig. 36c). Therefore, one can conclude that silicon atoms in silicene interact
with Ag(1 1 1) surface homogeneously, which is beneficial for continuous growth of high-quality silicene sheet.
Using DFT calculations, several groups have considered a variety of superstructures of monolayer silicene on Ag(1 1 1),
p p p p p p
including 4  4, 13  13, 2 3  2 3 and 7  7 phases [24,190,191,194,195,202,222–231]. All these first-principles
calculations show strong interaction between silicene and silver substrate, which has significant influence on the buckling
geometry and electronic states of silicene. The interaction strength between silicene and Ag substrate can be described by
the binding energy defined as
Eb ¼ ðESi  Esub  Et Þ=N; ð21Þ
where ESi is the energy of freestanding silicene sheet, Esub is the energy of Ag substrate, Et is the energy of silicene/Ag system,
N is the number of silicon atoms. From GGA calculations, the binding energies for various silicene superstructures on Ag
p p p p
(1 1 1) surface are 0.368 eV (4  4), 0.336 eV (2 3  2 3) and 0.332 eV ( 13  13), respectively [232]. The larger binding
energy of 4  4 superstructure may account for its most frequent emergence in experiments [19,21–23,180]. In addition,
p p
these three different reconstructed structures manifest distinct influences on the Raman spectra: the (2 3  2 3) and
p p
13  13 phases having a shift of the 2D peak (overtone of the D peak), while the 4  4 phase being affected by resonance
of the G-like peak [228].
For 4  4-silicene superstructure on Ag(1 1 1) surface, AIMD simulations at 500 K shows no topological defect ever gen-
erated up to 7.5 ps, indicating high thermal stability [24]. In contrast, a comparative AIMD simulation of silicene on Rh(1 1 1)
surface shows that the geometry structure of silicene is severely destroyed and an amorphous Si–Rh alloying surface is
58 J. Zhao et al. / Progress in Materials Science 83 (2016) 24–151

Ó Macmillan Publishers Limited 2016


Fig. 38. (a) Ground state structures and their formation energies (eV per atom) for SiN clusters (N = 6, 10, 13, 16, 19, 22, 24) in vacuum; (b) planar SiN
clusters composed of hexagonal rings (left) and the corresponding structures after relaxation in vacuum (right); (c) geometries and formation energies (eV
per Si atom) of 2D SiN clusters on Ag(1 1 1) surface. Reproduced with permission from Ref. [24].

formed after 2.3 ps at 500 K. The instability of silicene on Rh(1 1 1) surface originates from the large local energy difference
(0.306 eV) of Si atom on different sites of Rh surface. Therefore, the moderate and homogeneous interaction between Si
atoms and Ag atoms explains the superiority of Ag(1 1 1) surface for epitaxial growth of silicene.
The strong Si–Ag hybridization can be further visualized by the plot of electron density distribution of the p and p⁄ states
of silicene supported on Ag(1 1 1) [195] (see Fig. 37). Modifications of the p and p⁄ characteristics of freestanding silicene
states due to the existence of Ag substrate and distribution of electron density for these states on Ag substrate are found
in all three hybrid structures considered. Recently, the interaction between silicene and Ag surface was analyzed in terms
of spatial charge localization [229]. It was found that the top Si atoms carry a slight positive charge (0.13 |e| per atom), while
the first Ag plane is negatively charged, resulting in is an electrostatic interaction between the Si top atoms and the below-
lying Ag atoms. The charge transfer occurring at the silicene/Ag(1 1 1) interface as well as the intermixing between 3p Si and
5s Ag orbitals lead to the formation of a metallic interface [226,230].
Based on the first-principles results of silicene on Ag(1 1 1) and Ir(1 1 1), Kaltsas et al. [202,227] identified several key
points that are important for the formation of Si wetting films on metallic substrates: (1) The extra Si adatoms that arrive
at the surface during deposition can be incorporated in the honeycomb network and form bonds with underlying Ag atoms.
As a result, the corrugation profile changes, giving rise to various polymorphisms of silicene film. (2) There is a lower
coverage threshold for the formation of complete honeycomb films. There is also an upper limit in surface density, beyond
which the adatom incorporation scenario does not work and the growth of a second layer starts. (3) Different substrates have
unequal areal densities and interacting strengths with Si atoms. Therefore, certain Si monolayer honeycomb structures with
coverages that are below the upper threshold for deposition on one substrate, may not form on another substrate because
they surpass the upper coverage limit in the latter case.
Recently, the structure, stability and electronic properties of ultrathin atomic layers of Si (with thickness from monolayer
to quad-layer) on Ag(1 1 1) surfaces were systematically studied by Guo et al. using DFT calculations [220]. Several stable
p p p p
structures for the silicene on Ag(1 1 1) with periodicities of 3  3, 3  3, and 7  7 with respect to the 1  1 silicene
were considered. It was found that the 2  1 p-bonded chain structure on Ag(1 1 1) is most stable among all the bilayer,
tri-layer and quad-layer silicon structures on Ag(1 1 1), and the 7  7 dimer-adatom-stacking-fault (DAS) Si(1 1 1)-surface
structures are also stable on Ag surfaces. Therefore, relaxed or reconstructed Si(1 1 1)-surface structures on Ag(1 1 1) are more
stable than the silicene structures owning to their lower surface energies. The reason why the 2  1 p-bonded chain and/or
7  7 DAS structures have not been observed on Ag(1 1 1) in experiments was attributed to the synthesizing method, where
the structures are always produced via depositing the Si atoms on Ag(1 1 1). Moreover, the structural tristability and
p p
bistability for the 3  3 and 3  3 multilayer silicene on Ag(1 1 1) were found. The calculated transition barriers are
indicative of the flip-flop motion among/between the tristable/bistable structures at low temperature. The flip-flop
p p
motion between two of the three 3  3 structures produces the honeycomb STM structures observed in experiments
[20,175].
Based on the theoretical understanding of silicene–substrate interaction, it is thus necessary to explore the mechanism of
epitaxial growth of silicene on Ag(1 1 1) and other metal surfaces from atomistic point of view. Typically, the process for
epitaxial growth of silicene on metal surface involves three stages: (i) silicon atoms are deposited on metal surface; (ii)
J. Zhao et al. / Progress in Materials Science 83 (2016) 24–151 59

silicon monomers or dimers aggregate into many small silicon clusters, but most of them dissociate into monomers, dimers
and small pieces again; (iii) continuous growth of silicon clusters beyond nucleation size to cover the whole metal surface.
Among the three stages, the second one – nucleation of silicene, is crucial for understanding and controlling silicene
growth on metal surface. In this regard, Gao and Zhao [24] explored the geometries and energies of small silicon clusters
on Ag(1 1 1) surface to elucidate the early growth behavior of silicene. In vacuum, small silicon clusters are known to prefer
3D configurations composed of Si6, Si9, and Si10 subunits (Fig. 38a) [233–235]. For comparison, small planar SiN clusters
(N = 6, 10, 13, 16, 19, 22, 24) as aggregates of six-membered rings (6 MRs) cannot retain their planar structures in vacuum
and would transform into severely buckled 3D configurations upon relaxation (Fig. 38b). This is simply because silicon atoms
do not like sp2 hybridization in vacuum. Therefore, homogeneous nucleation in vacuum is impossible for synthesis of silicene
and a completely wetting substrate is needed.
To check the wetting behavior of Ag(1 1 1) surface in terms of silicon atoms, 6 MR-based clusters as small patches of
silicene clusters (Fig. 38b) as well as the ground state 3D configurations for gas-phase silicon clusters (Fig. 38a) were placed
on the Ag(1 1 1) surface, respectively. After relaxation, except for Si6 and Si13, the planar honeycomb structures of small
silicene patches can be preserved on Ag(1 1 1), which in turn can act as the nucleation center during the growth of silicene.
Due to metal passivation effect discussed above, the formation energies of clusters decrease after deposition on Ag(1 1 1)
surface. Interestingly, on Ag(1 1 1) surface, the planar SiN clusters have lower formation energies than the corresponding
ground state SiN cluster in vacuum. In the other words, silicon clusters prefer 2D planar structures to 3D structures on Ag
(1 1 1) surface, which confirms that Ag(1 1 1) surface is a substrate of wetting. Moreover, the formation energy for
silicene-like 2D cluster decreases as the number of hexagonal ring increasing, suggesting that aggregation of silicon atoms
is energetically favorable. In a recent experiment [183], small silicon clusters (1.2 nm in diameter) as the nuclei of silicene

Ó IOP Publishing 2016

Fig. 39. Experimental valence band electronic structure of 3  3 silicene film on (1 1 1) Ag surfaces at the other C point of the 3  3 supercell Brillouin zone.
(a) ARPES spectrum measured at the C02 point, along the M—C—M direction of the 3  3 lattice. Empty yellow dots represent the unaffected ‘sp’ silver
substrate band. (b) A discrete representation of the ARPES spectra measured at the C00 (blue dots) and at the C02 (blue crosses) points along the M—C—M
direction of the 3  3 lattice. Black lines represent the calculated energy bands of the 3  3 silicene/Ag(1 1 1) surfaces in the supercell Brillouin zone and
empty black dots are the same as the yellow dots of panel (a). MDC’s taken at 96 eV, 112 eV and 130 eV photon energies are presented in panels (c and d) for
clean Ag(1 1 1) and the 3  3 silicene phase, respectively. The MDC’s have been recorded at the M Ag symmetry point of the Ag 1  1 BZ at a constant binding
energy of 1.0 eV. Reproduced with permission from Ref. [198].
60 J. Zhao et al. / Progress in Materials Science 83 (2016) 24–151

growth were recognized as dotted spots in the STM topography of 0.77 ML Si on Ag(1 1 1), in agreement with the theoretical
prediction.
Another important factor for the growth of silicene is the diffusion barrier for silicon atom on Ag(1 1 1) surface. As shown
in Fig. 38c, the most stable adsorption site for silicon monomer is the hollow site. Nudged elastic band (NEB) calculation
showed that the diffusion barrier for a silicon monomer on Ag(1 1 1) is as low as 0.031 eV, implying easy migration of silicon
atom on Ag surface [24]. After being deposited on Ag(1 1 1) surface, the silicon atoms would quickly aggregate into small
planar clusters by overcoming a small barrier. Then small patches of silicene will occur at some critical cluster sizes and grow
into large silicene sheet by continuously absorbing silicon atoms on the Ag surface.
A successive theoretical study by Shu et al. demonstrated that the nucleation rate of silicene is very sensitive to both
growth temperature and chemical potential Dl of the silicon atoms [236]. At a typical growth temperature (500 K), the
nucleation rate shows a 23 orders of magnitude increase (from 1010 to 1013 cm2 s1) as Dl alters from 0.05 to
0.15 eV. For a specified Dl (e.g. Dl = 0.1 eV), reduction of the temperature from 500 K to 300 K would lead to an 11 orders
of magnitude drop in the nucleation rate (from 107 to 104 cm2 s1). Usually, a high nucleation rate would result in large
amounts of defects. Thus the growth conditions should be carefully chosen in order to reduce the defects in silicene.
During the third stage for epitaxial growth of silicene, various silicene superstructures can form on Ag(1 1 1) surface,
depending on the dose of deposited silicon atoms and the substrate temperature. Based on micro-LEED results, Acun
et al. argued that silicene layers are intrinsically unstable against the formation of an ‘‘sp3-like” hybridized, bulk-like silicon
structure [186]. The stability and structural properties of several Si monolayer films on a Ag(1 1 1) surface were investigated
by Kaltsas and co-workers through DFT calculations [202]. They proposed that compact honeycomb sheets of silicon atoms
on Ag(1 1 1) surface are favorable as the dose of silicon atoms on Ag substrate increases. Transformation from low-coverage
structures to denser monolayer configurations may occur either through coalescence of the former, or through incorporation
of Si adatoms in the honeycomb network [202]. When the coverage of silicene is more than one monolayer (ML), a second
layer emerges on the first layer of silicene in the form of an island [178]. These islands prefer to be located on the terrace
edges of the first layer. STM observation in experiments shows that there is no trace of continuity between the first layer
and the second layer, indicating that the second layer grows when the first layer film has been completely formed on the
silver substrate. Based on STM observation, Xu and co-workers proposed that the silicene sheets prefer to grow from the
terrace edge of Ag(1 1 1) substrate following the layer-plus-island growth mode, and that the critical layer for the transition
from 2D to 3D mode is 1 ML [178].
The substrate temperature also plays an important role in the formation of silicene with different structures [178]. When
p p
the substrate temperature is less than 450 K, an amorphous silicon film forms on the Ag(1 1 1) substrate. 13  13 and
4  4 phases coexist on a large area of the silver surface when the substrate temperature is between 450 K and 520 K. When
p p
the substrate temperature is increased above 520 K, the 2 3  2 3 phase forms, although there are traces of 4  4 phase.
p p
Pure 2 3  2 3 phase silicene can be obtained when the substrate temperature is higher than 550 K. Using LEED and Auger
electron spectroscopy (AES), Rahman et al. [237] monitored the surface structure evolution of Si adsorption on Ag(1 1 1).
With varying substrate temperature during deposition, silicon atoms form pure phase of 4  4 silicene on Ag(1 1 1) at
520 K, but form 2D surface Si–Ag alloy at temperature higher than 620 K. Very recently, Grazianetti emphasized the critical
role of post-deposition annealing (PDA) temperature in silicene growth as an additional parameter in controlling the silicene
synthesis [183]. In the submonolayer coverage regime, PDA induces the self-organization of silicene islands; while for mono-
p p p p
layer coverage, phase transitions among the three main silicene superstructures (i.e., 4  4, 13  13, and 2 3  2 3) can
be induced by controlling the PDA temperature.

3.1.7. Spectroscopic characterization and electronic structures


Following the successful preparation of silicene, the focusing issue is whether silicene indeed exhibits the exotic
electronic properties (especially, the Dirac electronic state) predicted in theory. So far, ARPES and scanning tunneling micro-
scopy/spectroscopy (STM/STS) are the two major spectroscopic techniques to characterize the electronic structures of
silicene. Le Lay and co-workers reported, together with their silicene structure, possible existence of the Dirac cone and
linear dispersion in the 4  4 phase of silicene from ARPES measurements [19]. Later, Avila et al. [198] revisited the elec-
tronic structures of the 3  3 silicene (4  4 using the present notion) on Ag(1 1 1) surface by using a compiled set of LEED,
high-energy resolution core levels (CLs) and ARPES data. As displayed in Fig. 39a and b, ARPES spectra demonstrated exis-
tence of a silicene-derived band at several C points of the 2D Brillouin zones. The spectra highlight two linearly dispersing
bands, which cross each other at 0.3 eV below the Fermi level and have an estimated Fermi velocity of 1.3  106 m/s.
Analysis of the momentum distribution curves (MDC) taken at different photon energies of the bands around the C33 points
(Fig. 39c and d) further confirmed that the linear dispersion comes from silicene rather than Ag substrate.
Soon after the first ARPES experiment reported for 4  4 silicene on Ag(1 1 1) [19], there has been a fierce debate about
whether the Dirac cone of monolayer silicene can be preserved on Ag(1 1 1) surface [19,190,191,194–196,198,199,221–22
5]. The key issue is that, once a monolayer silicene sheet is placed on a clean Ag(1 1 1) substrate, the strong interaction
between the silicene film and the Ag substrate may significantly influence the surface electronic band structure.
Combining spectroscopic signatures of quantum Hall effect by the Landau quantization under a magnetic field and DFT
calculations, Lin et al. [190] proposed that silicene would no longer be a 2D Dirac fermion system when synthesized on the
Ag(1 1 1) surface due to substrate-induced symmetry breaking. Fig. 40a shows the STM image of a (4  4) superstructure, and
the amplificatory STM image is given in Fig. 40b. STS spectra were used to measure the Landau levels. For the magnetic field
J. Zhao et al. / Progress in Materials Science 83 (2016) 24–151 61

Ó American Physical Society 2016

Fig. 40. (a) STM image of large area of silicene (sample bias voltage VS = 0.70 V and tunneling current It = 0.19 nA. The image size is 35  35 nm2). (b) High
resolution STM image of the 4  4 silicene (VS = +0.50 V and It = 0.30 nA, 3.65  3.65 nm2). The unit cell is shown by the white rhombus. (c) STS spectra of
silicene for various magnetic fields perpendicular to the sample surface, BZ. (d) STS spectra of HOPG for various BZ. The purple triangles and green circles
show the peaks originating from the Landau levels of massless and massive Dirac fermions, respectively. The Landau level with n = 0 is marked by the
yellow bar and the Landau level with n = 1 of massive Dirac fermions is not clearly resolved in 3 T due to low magnetic field. Reproduced with permission
from Ref. [190].

perpendicular to the sample between zero and 7 T, characteristic structures due to the Landau quantization were not found
(Fig. 40c), which implies the silicene loses its characteristic of Dirac fermion system. In contrast, Landau levels appear in the
highly-oriented pyrolytic graphite (HOPG) (Fig. 40d), which can be reasonably rationalized by a combination of massless and
massive Dirac fermions.
Using the complementary techniques of DFT and soft X-ray spectroscopy at the Si L2,3 edge, Johnson et al. [199] examined
the atomic and electronic structures of five distinct epitaxial silicene morphologies on Ag(1 1 1). They found that strong
sp3-like hybridization between Si atoms and the Ag substrate’s d states as well as the presence of the Ag sp band causes
the Si pz states to span the Fermi level. As a consequence, epitaxial silicene on Ag(1 1 1) is strongly metallic and does not
possess the Dirac cone electronic structure.
62 J. Zhao et al. / Progress in Materials Science 83 (2016) 24–151

Ó American Physical Society 2016


Fig. 41. Left panel: bands of reconstructed (3  3) silicene (4  4 using the present notion) in the absence of Ag substrate (unsupported silicene) unfolded to
BZ of (1  1) silicene are shown by red dots. The radii of dots correspond to the weight of unfolding. The band structure of ideally buckled silicene is shown
by green lines. Right panel: unfolded band structure of silicene on Ag. Red dots correspond to states with significant contribution from silicene. Reproduced
with permission from Ref. [194].

DFT calculations by different groups also support that the Dirac cone of silicene is destroyed on Ag(1 1 1) surface due to
strong interaction and mixing between Si and Ag orbitals. It was suggested that the linear dispersions observed in experi-
ments originate either from the Ag substrate or from the hybridization states between silicene and Ag(1 1 1) surface [190,
191,193–196,203,221–225]. For instance, Guo et al. [191,203] considered several silicene structures with 4  4,
p p p p
13  13, 2 3  2 3 periodicities with respect to the (1  1) Ag(1 1 1) lateral cell and recognized two factors that account
for the disappearance of Dirac electrons: (1) the substantial buckling of silicene layer leading to the p state rehybridization
with the r state and consequent the p+ state; (2) the mixing of the p+ state with the states of the Ag atoms in the substrate,
converting the p+ state to the mixed p+ state and thus making the state shift downwards or upwards. From the projected
effective band structure, Wang and Cheng [221] suggested that the linear dispersions observed in ARPES experiments come
from the sp band of Ag(1 1 1) surface instead of from silicene. Similarly, Mahatha et al. [224] argued that previously assigned
p-silicene state actually derives from an interface Ag state of free-electron-like character.
As shown in Fig. 41, the reconstructed (3  3) silicene (4  4 using the present notion) without Ag substrate opens a
0.3 eV band gap at the pristine Dirac cone (red dots in the left panel) due to structural modification and symmetry breaking
induced by the Ag substrate [191,194,222,226,230]. Comparing the unfolded band of reconstructed (3  3) silicene (red dots
in the left panel) with the unfolded band of silicene/Ag (red dots in the right panel), no one to one correspondence can be
found between them, indicating a strong hybridization between silicene and Ag substrate.
In most DFT calculations, the Ag(1 1 1) substrate was simulated by finite-thickness slab models with only a few layers of
Ag atoms. To overcome the limitation, Ishida et al. [223] investigated the electronic structure of 4  4 silicene monolayer on
a semi-infinite Ag(1 1 1) substrate using DFT and embedded Green’s function technique. The results reconfirmed that the 2D
Dirac bands do not exist on this surface owing to the symmetry breaking and strong orbital hybridizations between the Si p
and Ag sp states. Taking the advantage of the semi-infinite calculation, the details of the silicene-induced electronic states,
especially their spectral shape as a function of energy at each k were examined. A possible mechanism for the 2D
free-electron-like band to emerge at the interface due to strong Ag–Si interaction was suggested.
The debates on this issue still continue, although at present most researchers tend to believe that there is no Dirac state in
this system. Recently, however, Wu and Zhou’s group has directly observed Dirac cones on a surface consisting 4  4 silicene
monolayer on Ag(1 1 1) by ARPES [197]. Interestingly, the Dirac cones are present at the edges of the BZ, and there are twelve
J. Zhao et al. / Progress in Materials Science 83 (2016) 24–151 63

Ó American Physical Society 2016

Fig. 42. (a) dI/dV curves taken at 77 K. The position of Dirac point (DP) is labeled. (b) The STM image (40  40 nm2) of 1 ML silicene surface containing an
island of second layer taken at tip bias 1.0 V. (c, d and e) dI/dV maps of the same area as (b) taken at tip bias 1.0 V, 0.8 V and 0.5 V, respectively.
(f) Energy dispersion for silicene determined from wavelength of QPI patterns. The inset shows a schematic drawing of the overall band structure, with the
relative location of DP, EF and our data points (red line). Reproduced with permission from Ref. [175].

Dirac cones. This is in contrast to the theoretical expectation for freestanding silicene that involves six Dirac cones at the K
(K0 ) points of the Brillouin zone. The unusual Dirac cone structure is a consequence of the silicene–Ag(1 1 1) interface, and
such interfacial effects may open up new possibilities for investigating quantum phenomena in low dimensional structures.
In contrast to the controversy about the electronic states of 4  4 silicene on Ag(1 1 1), low-temperature electronic state
p p
measurements by STS conducted by Wu’s group have revealed that the 3  3 phase exhibits a metallic surface state
which generates quasi-particle interference (QPI) patterns [175]. As shown in Fig. 42, the fitting of energy–momentum
relation of the QPI pattern around a step edge gives linear dispersion that is expected for a Dirac fermion system. The same
group has also measured the decaying behavior of the QPI patterns, and obtained a large decaying factor, which is typical for
backscattering suppressed systems like topological insulator or graphene, and different from conventional 2D gas systems
64 J. Zhao et al. / Progress in Materials Science 83 (2016) 24–151

Ó The Electrochemical Society 2016


Fig. 43. (a) Experimental Raman spectrum of Al2O3-capped silicene acquired with the 488 nm laser line. (b and c) show computed Raman spectra of the
p p
13  13 and 4  4 superstructures, respectively, obtained by the calculated vibrational spectra convoluted with a uniform Gaussian broadening having a
p p
full width at half-maximum of 10 cm1. Schematic picture of the relevant Raman modes for the 13  13 superstructure (d) and for the 4  4
superstructure (e). Reproduced with permission from Ref. [239].

Ó American Chemical Society 2016

p p
Fig. 44. (a) STM topographic image (U = 1.45 V, and I = 0.25 nA), showing a ( 7  7) superstructure of the silicon adlayer formed on the Ir(1 1 1) surface.
The direction of this reconstruction is indicated by the yellow arrow. The close-packed direction of Ir½1 1  0 is indicated by the white arrow. The angle
between the yellow and white arrows is around 41°. (b) STM image (U = 1.5 V, and I = 0.05 nA) of the silicon superstructure with another orientation,
indicated by the blue arrow. The rotation angle between the blue and white arrows is about 19°. (c) Line profile along the black line in b, revealing the
periodicity of the protrusions (0.72 nm) and a corrugation of around 0.6 Å for the silicon ad-layer. (d) Zoomed-in STM image of the silicon layer. Besides the
brightest protrusions, two other regions showing different contrast are indicated by the upward and downward triangles. The honeycomb feature is
indicated by the black hexagon. (e) Simulated STM image, showing features identical with the experimental results in the same triangles and hexagons.
p p p p
(f) Top view of the relaxed atomic model of the ( 3  3)silicene/( 7  7)Ir(1 1 1) configuration. Reproduced with permission from Ref.
[29].

p p
[238]. This provides further evidences that the 3  3 phase host Dirac fermions. For comparison, although other mono-
layer silicene phases are also expected to be metallic, QPI patterns have never been observed on their surfaces. Moreover, a
p p
number of intriguing phenomena were also observed in the 3  3 silicene, including the low temperature phase
J. Zhao et al. / Progress in Materials Science 83 (2016) 24–151 65

transition [218], and a superconducting like gap at zero bias [44]. These results demonstrated that silicene is a promising
playing ground for exotic electronic properties.
In addition to the above ARPES and STS results, the atomic structures, chemical bonding, vibrational properties and
electronic states of epitaxial silicene films on Ag(1 1 1) have been extensively characterized by other spectroscopic means,
such as X-ray photoemission spectroscopy (XPS) [23,183,239], soft X-ray spectroscopy [199], reflection high-energy positron
diffraction [240], Raman [42,181,182,239,241], and extended X-ray absorption fine structure (EXAFS) [216]. For instance,
Chiappe et al. [23] deposited ultrathin silicon layers on the Ag(1 1 1) surface with coverage of 0.45 ML and 0.65 ML, respec-
tively. The measured XPS features of the Si 2p and Ag 4s core levels showed a marginal chemical interaction between the Ag
substrate and the Si ad-layers, ruling out the possible formation of Si–Ag alloy.
Raman spectra of various silicene structures on Ag(1 1 1) substrates have been obtained by different groups
[42,181,182,239]. Molle and co-workers [181,239] performed ex-situ Raman characterization of several silicene superstruc-
tures epitaxially grown on Ag(1 1 1) substrates and capped by a 4 nm Al2O3 ad-layer. The Raman spectra of the encapsulated
silicene sheets have been interpreted by comparing with the DFT simulated spectra. From DFT calculations, the double
p p
degenerate E2g mode peaked at 505 cm1 and 495 cm1 (Fig. 43b and c) are obtained for the 13  13 and 4  4 super-
1
structures, respectively, which are in good agreement with the experimental feature at 516 cm (Fig. 43a) and can be
related to the G-like peak at 575 cm1 calculated for freestanding silicene with low buckling [242]. As illustrated in
Fig. 43d and e, this peak comes from the bond stretching of all pairs of sp2 silicon atoms in the silicene unit cell, thus
representing an unambiguous fingerprint of a hexagonal 2D lattice. Unlike the case of freestanding low-buckled silicene with
only one peak [242], other Raman active modes are found in the calculated spectra as D, T and K in Fig. 43b and c. These
vibrational modes are peculiar for each silicene superstructure since they reflect the intrinsic ‘‘disorder” of silicon atoms
related to the non-uniform substrate-induced buckling. Specifically, for the 4  4 case, the D(A1g) mode originates from
the breathing-like displacement of planar hexagons, while the T(A1g) is related to the breathing-like displacement of non-
p p
planar hexagons (Fig. 43e). The D(A1g) and K(B2u) modes for the 13  13 phase arise from the breathing mode of the
hexagonal rings having alternating up- and down-standing atoms and to Kekule distorted 30 hexagonal rings, respectively
p p
(Fig. 43d). Moreover, the Raman spectrum of 2 3  2 3 silicene is characterized by a strong E2g mode at 521 cm1 along
1
with some weak A1g modes between 300 and 460 cm , which indicates a lower amount of intrinsic disorder with respect
p p
to the 13  13 and 4  4 superstructures.

Ó American Physical Society 2016

Fig. 45. STM images of the (2  2)-reconstructed ZrB2(0 0 0 1) surface with different length scales: (a) 20  9.5 nm2, I = 55 pA, Vs = 700 mV, (b) 4.2  2 nm2,
I = 600 pA, Vs = 100 mV. The white lines emphasize the direction of offsets between successive domains. The (2  2) unit cell and the honeycomb mesh are
emphasized by green and blue solid lines, respectively. (c) Calculated structure of epitaxial silicene on Zr-terminated ZrB2(0 0 0 1) showing height profile. (d)
Brillouin zones of the (2  2)-reconstructed ZrB2(0 0 0 1) surface and of an unreconstructed silicene layer. (e) ARUPS spectra along the C—K—M and C—M—C
high symmetry direction. Features Sn and Xn relate to surface electronic states of diboride and those of epitaxial Si layer, respectively. Reproduced with
permission from Ref. [27].
66 J. Zhao et al. / Progress in Materials Science 83 (2016) 24–151

Recently, Du et al. [182] reported in-situ Raman spectroscopy of epitaxial silicene on Ag(1 1 1) surface with different
p p
structures and coverage. Generally, the Raman spectra from in-situ measurement of the 13  13/4  4 structures of
silicene are consistent with the ex-situ results by Molle et al. [181,239], that is, the E2g mode around 530 cm1 as the major
peak plus five Raman peaks in the range of 200–500 cm1 (marked as D1–D5) in the low-coverage samples. By carefully
analyzing the shift of Raman peaks, they concluded that electron–phonon coupling in silicene is significantly modulated
by tensile strain due to the lattice mismatch between silicene and substrate, as well as the charge doping from the substrate.

3.2. Silicene monolayer on other substrates

Besides the most frequently used Ag(1 1 1), silicene monolayer sheets have been also fabricated on a few other substrates,
including ZrB2(0 0 0 1) [27,243–245], ZrC(1 1 1) [28], Ir(1 1 1) [29], and MoS2 [30]. Fabrications of silicene on substrates other
than Ag(1 1 1) suggest that there are many opportunities of finding suitable growth substrates for silicene, and perhaps, for
germanene and stanene.
Soon after the synthesis of monolayer silicene on Ag(1 1 1), Gao’s group [29] deposited silicon atoms on the Ir(1 1 1)
surface at room temperature and then annealed the sample to 670 K. As shown in Fig. 44, a continuous atomic layer with
p p p p
a buckled conformation was obtained, which was a ( 3  3)silicene/( 7  7)Ir(1 1 1) superstructure as characterized
by LEED, STM and confirmed by DFT calculations. Analysis of electron localization function demonstrated a strong Si–Si
covalent bonding throughout the silicene layer.
The silicene–substrate interaction and electronic structure of silicene monolayer on Ir(1 1 1) have been further investigated
p p
by DFT calculations [196,246]. Silicene forms ( 3  3) superstructure on Ir(1 1 1) surface with a buckled height of 0.83 Å and
the binding energy between silicene and Ir(1 1 1) surface is 1.69 eV [196], much larger than that between silicene and Ag
(1 1 1), i.e., around 0.4 eV (see Section 3.1.6). This indicates the a strong adsorption of silicene on Ir surface, which
originates from strong hybridization between 3pz state of Si and 5d states of Ir through projected density of states (PDOS)
analysis [246]. Consequently, silicene on Ir(1 1 1) loses its original Dirac cone around the Fermi level, leading to a metallic band
structure. Recently, Dai’s group [246] has performed systematical DFT calculations on the atomic structures, growth behavior,
surface phase diagram, electronic band structure of silicene on Ir(1 1 1) substrate. They found that the experimentally
p p
observed 3  3 superstructure is thermodynamically most stable with increasing the temperature toward a silicon chem-
ical potential range of 5.873 to 5.255 eV. Moreover, a hidden linear-dispersive band was predicted for silicene on Ir(1 1 1).
However, combining DFT calculations and a collection of in situ experimental measurements including STM, dynamic
atomic force microscopy, LEED, synchrotron radiation photoemission spectroscopy, and ARUPS, Švec et al. found that silicon
atoms would not form a silicene structure on Pt(1 1 1) surface, and would favor the formation of Si-Pt surface alloy consisting

Ó American Chemical Society 2016

Fig. 46. Optimized structure model of Si on ZrC(1 1 1): (a) Bird’s eye, (b) planar, and (c) side views. (d) Specular HREELS (E0 = 9 eV) of the Si on ZrC(1 1 1) (red
line) and calculated PDOS of the vertical vibrations of Si at C with 1 meV Gaussian broadening. (e) Calculated valence charge density for ZrC(1 1 1)2  2-Si.
 section including SiA and SiB. Reproduced with permission from Ref. [28].
The front face is a ð1 1 1Þ
J. Zhao et al. / Progress in Materials Science 83 (2016) 24–151 67

Fig. 47. STM based identification of the epitaxial Si nanosheet on MoS2. (a) STM image of a clean MoS2 surface (VB = 1 V, IS = 0.7 nA). (b) STM image of a
partially covered MoS2 surface after deposition of 0.8 ML of Si (VB = 1 V, IS = 0.7 nA). The blue contour delimits an uncovered portion of the MoS2 substrate.
Inset: the RHEED pattern exhibits well-defined characteristic streaks. (c) Higher resolution STM image (VB = 0.2 V, IS = 2 nA) of a partially covered surface.
The left side and the right side of the image correspond to a portion of bare MoS2 and a Si covered region, respectively. A line profile taken across the two
terraces allows measurement of the amplitude of the step, which amounts to about 5 Å. In (d), the magnified topography (right inset) shows a hexagonal
surface pattern whose periodicity is extracted from the analysis of the self-correlation function. A sketch (bottom inset) replicates the Si honeycomb
structure. Reproduced with permission from Ref. [30]. Copyright 1999–2016, John Wiley & Sons, Inc.

of Si3Pt tetramers instead [247]. Based on DFT calculations, they further argued that a Si–Ir surface alloy rather than silicene
layer would form when Si atoms are deposited on Ir(1 1 1) surface [247].
Parallel to the synthesis of monolayer silicene on Ag(1 1 1), Fleurence and co-workers demonstrated that epitaxial 2D
silicene forms spontaneously through surface segregation on zirconium diboride thin films grown on Si(1 1 1) wafers [27].
The oxide-free single-crystalline ZrB2 thin film surface grown on Si(1 1 1) exhibits a (2  2) reconstruction, which is related
to Si adatoms segregating from the Si substrate. The presence of Si adatoms at the surface was confirmed by photoelectron
spectrum of Si 2p states [27]. The large-scale STM image of the (2  2)-reconstructed ZrB2(0 0 0 1) surface shown in Fig. 45a
exhibits a stripe pattern. The driving force behind the mechanism that triggers the appearance of the stripe pattern was
found to be associated with a phonon instability characterized by a zero-frequency mode reaching a critical point of
instability [248]. Fig. 45b reveals fine details of the surface, which correspond to a honeycomb mesh with the lattice constant
of approximately 3.65 Å. A metastable structure of silicene layer on ZrB2(0 0 0 1) obtained by first-principles calculations
(depicted in Fig. 45c) is consistent with the structural information from experiments. In this structure, there are three
different silicon atoms labeled by SiA, SiB and SiC (Fig. 45c). The calculated SiA–SiB and SiB–SiC bond lengths are 2.266 Å
and 2.242 Å, respectively, which are very close to the predicted bond length for freestanding silicene [8].
This structure for silicene on ZrB2(0 0 0 1) was further confirmed by comparison of the measured STS and the computed
local density of states [243]. The ratio of Si atoms at various sites was studied using high-resolution Si 2p photoelectron spec-
p p
tra [244]. The Brillouin zones of the (2  2)-reconstructed ZrB2(0 0 0 1) surface, which corresponds to the ( 3  3)-
reconstructed silicene, and the unreconstructed silicene are presented in Fig. 45d. Angle-resolved ultraviolet photoelectron
spectroscopy (ARUPS) spectra along the C—K—M and C—M—C directions of the (2  2)-reconstructed ZrB2(0 0 0 1) surface as
a function of the in-plane electron wave number k|| are shown in Fig. 45e in form of intensity plots. The upward curvature of
X2 bears some resemblance to the predicted Dirac cone of p bands of freestanding silicene at KSi [8], proving that the bonding
within the Si layer has a large degree of sp2 character. By means of low-temperature STM and STS, Fleurence et al. also
demonstrated the n-type semiconducting nature of epitaxial silicene on ZrB2(0 0 0 1) [243]. Recently, the phonon dispersion
of silicene on a ZrB2(0 0 0 1) surface was measured using high resolution electron energy loss spectroscopy (HREELS) and
further illustrated with the aid of first-principles calculations [245].
Since the epitaxial silicene is not chemically inert under ambient conditions, an insulating capping layer is needed for its
device applications or ex-situ characterization outside of the UHV environments. The feasibility of depositing an aluminum
68 J. Zhao et al. / Progress in Materials Science 83 (2016) 24–151

Ó Macmillan Publishers Limited 2016


Fig. 48. Band structures of epitaxial silicene on various metal surfaces. The Dirac point is folded at the C point. The Fermi level is set to zero. The red color in
(a–b) and (d–f) indicates the states contributed by the Si atoms, and the green, blue, and magenta colors in (c) indicate the states contributed by s, px and py,
and pz orbitals of the Si atoms, respectively. The thickness of these colors is proportional to the Si atom character. The inset in (f) is the electron density at
the C point inside the black square, and the yellow and silver balls are Si and Cu atoms, respectively. Reproduced with permission from Ref.
[196].

nitride (AlN) capping layer on top of epitaxial silicene on ZrB2 thin films on Si(1 1 1) substrates, using trimethylaluminum
(TMA) and ammonia (NH3) as precursors, was investigated by Van Bui et al. [249]. They found that an AlN-related layer
can indeed be grown. However, the silicene layer reacts strongly with both precursor molecules, resulting in the formation
of Si–C and Si–N bonds such that the use of these precursors does not allow for the protective AlN encapsulation. On the
other hand, different from the structure model shown in Fig. 45c which is consistent with the experimental data, a
p p
planar-like, ( 3  3)-reconstructed silicene sheet was calculated to become the most stable epitaxial structure on
ZrB2(0 0 0 1) [250]. Also note that strain relief at the boundaries of stress domains (shown in Fig. 45a) may reduce the ampli-
tude of buckling and affect the relative stability of possible structures [27].
Similar to ZrB2(0 0 0 1), ZrC(1 1 1) is also utilized as a substrate for silicene growth. In the experiment by Aizawa et al. [28],
Si atoms were deposited on ZrC(1 1 1) film grown epitaxially on an NbC(1 1 1) single crystal surface. The as-fabricated silicene
sheet was then characterized by phonon dispersion measurements using high-resolution electron energy loss spectroscopy
(HREELS), combined with DFT calculations. It was confirmed that a distorted silicene monolayer is formed on the
p p
Zr-terminated ZrC(1 1 1), in which the 3  3 unit cell of silicene lattice coincides with the 2  2 unit of the substrate
ZrC(1 1 1), see Fig. 46a–c. The lattice dynamics from DFT simulations matched well with the measured phonons, as displayed
in Fig. 46d. The calculated valence charge density in Fig. 46e shows weak covalent bonding between the SiB atom and the
subsurface Zr atom.
Similar to the case of silicene on Ag(1 1 1), strong interaction between silicene and ZrB2(0 0 0 1) or ZrC(1 1 1) surfaces is
revealed by DFT calculations and thus silicene loses its Dirac cone when it is synthesized on these two substrates
[28,250,251]. On ZrB2(0 0 0 1) surface, all silicene-derived bands are hybridized with Zr d electronic states to some extent.
Although the upward curved bands in the vicinity of the Fermi level at the KSi(1  1) point are of partial p character, they
are hybridized with Si s, px, py, and Zr d orbitals, reflecting the intermediate sp2/sp3 hybridization of the epitaxial silicene
J. Zhao et al. / Progress in Materials Science 83 (2016) 24–151 69

Ó American Chemical Society 2016


Fig. 49. Calculated silicene structures formed on Al(1 1 1): (a) honeycomb silicene, (b) polygonal silicene. The yellow spheres represent Si atoms, and the
blue spheres represent Al atoms. The upper and lower panels show the top and side views, respectively. (c) Schematic configuration of the polygonal
silicene showing its tessellation. The unit cell for the polygonal layer is shown in red. Reproduced with permission from Ref. [254].

[251]. In contrast to the experimental STM observation of regularly buckled geometry, a planar-like phase was found as the
ground state structure of silicene on the Zr-terminated ZrB2(0 0 0 1) surface, which is stabilized by interactions of every Si
atoms with Zr d orbitals [250]. This suggests a richer phase diagram of epitaxial silicene, as compared with freestanding
silicene.
Using MBE technique, 2D Si nanosheets have been successfully fabricated on MoS2, which is a semiconducting substrate
[30]. STM image of a clean MoS2 surface is given in Fig. 47a, which shows a periodic array of hexagonally ordered bright
spots. Then a Si layer with nominal coverage of 0.8 ML was deposited on the clean MoS2 crystal at a substrate temperature
of 200 °C. The STM image in Fig. 47b evidences the presence of a Si nanosheet locally covering the MoS2 surface. An ordered
surface pattern can be observed from the Si covered regions in Fig. 47c. The relative heights of atoms along the line profile
taken across the two terraces of Si nonosheet and MoS2 surface (Fig. 47c) were measured, which yield a buckled height of
2 Å in the Si nanosheet and a interlayer distance of 3 Å between the Si nanosheet and MoS2 substrate. Hence, the derived
Si nanosheet corresponds to a high-buckled silicene monolayer. The magnified topography of the 2D Si domain (Fig. 47d)
shows a clear hexagonal surface pattern, suggesting that Si atoms locally self-organize in honeycomb lattice. The lattice
constant of the high-buckled silicene layer is 3.2 Å, which can be extracted from the analysis of the self-correlation function
in Fig. 47d. Both STS and DFT calculations indicate that the obtained high-buckled silicene is metallic [30]. Although the
silicene sheet on MoS2 surface is severely distorted from the freestanding one, this is the first step towards synthesis of
silicene on nonmetal substrates, which is the key towards future nanoelectronic devices based on silicene.
Besides those successful substrates realized in experiments, the interactions between silicene and many metal substrates,
such as Al, Mg, Au, Cu, Pt, Pb, Ca and Ca-functionalized Si(1 1 1) surface, have been extensively explored by first-principles
calculations [196,252–254], aiming to find other possible substrates for growth of silicene and to elucidate the influence
of metal substrate on the electronic structure of silicene. According to the calculated binding energy defined by Eq. (21),
the strength of interaction between silicene and metal substrates ascends in the following order: Al < Mg < Ag < Au <
Cu < Ir < Pt [196]. The band structures of epitaxial silicene on some metal surfaces are shown in Fig. 48, where the red dots
stand for the bands contributed by Si atoms. One can hardly see the original ‘‘cone” shape band structure of freestanding
silicene in all the systems examined, suggesting a strong band hybridization between silicene and metal substrates. There-
fore, the absence of the Dirac cone is a common feature for epitaxial silicene on metal substrates, including at least Ir, Cu, Mg,
Au, Pt, Al, and Ag [196]. However, intercalation of potassium (K) atoms between silicene and these metal substrates would
recover the Dirac cone of silicene [196], similar to the case of K intercalation in bilayer silicene [35]. The recovered Dirac cone
is located at 0.40–0.78 eV below EF, suggesting an n-type doping of silicene. Specially, for silicene on Ag, Mg Cu and Ir
substrates, K intercalation would open a band gap of 0.15–0.40 eV at the original Dirac point, while for silicene on Pt, Al
and Au substrates, no band gap would be opened after K intercalation. Instead of directly placing silicene on Cu(1 1 1) surface,
introducing a hexagonal boron nitride (h-BN) buffer layer between silicene and Cu(1 1 1) can block the strong interaction
between silicene and Cu(1 1 1) and thus preserve the Dirac point in the original silicene, but with the Dirac cone about
0.2 eV below the Fermi level due to the charge transfer from substrate to silicene [255].
In analogy to the 4  4 silicene superstructure on Ag(1 1 1) surface, Morishita et al. [254] predicted that the 4  4 honey-
comb structure of silicene can be stably formed on the Al(1 1 1) surface (Fig. 49a) using AIMD calculations. In addition,
silicene on Al(1 1 1) may adopt a new ordered monolayer structure consisting of triangles, rectangles, pentagons, and hexa-
gons, namely, ‘‘polygonal silicene” (Fig. 49b and c), which is less stable than the regular honeycomb silicene by only
70 J. Zhao et al. / Progress in Materials Science 83 (2016) 24–151

0.0135 eV/atom. Moreover, the polygonal silicene exhibits a free electron-like density of state with high electronic conduc-
tion and is suitable as electrode materials. Though this theoretical prediction has not been fulfilled in experiment yet, Al
(1 1 1) has been used as a substrate for successful synthesis of germanene monolayer [256,257].
Recent DFT calculations showed that several superstructures of silicene may form on Pb(1 1 1) surface, including
p p p p
( 7  7)R19.1°, ( 13  13)R13.9° and (4  4) with respect to silicene [252]. The binding energies between silicene
and Pb surface ranges from 0.15 to 0.18 eV per Si atom, which are lower than most other metal surfaces. Benefited from
the relatively weak interaction, the linear dispersion of silicene bands around the K points of the BZ survives and the elec-
tronic bands have mainly 3p character of silicene with very little contribution of the 6p states of Pb [252]. Ca(1 1 1) surface
was also proposed as a possible substrate for silicene growth [253]. In contrast to Mg supported silicene, the Dirac cone of
silicene would not be destroyed by Ca(1 1 1) surface. Instead, a small energy gap of 0.55 eV would be opened at the Dirac
point, which locates below the Fermi level. Besides, Ca-functionalized Si(1 1 1) surface, i.e., Ca(1.0 ML)/Si(1 1 1), would also
preserve the Dirac cone of silicene, similar to that of Ca(1 1 1) surface [253].
There have also been some theoretical efforts to searching suitable nonmetal substrate for growth of silicene [258–260].
For instance, Kokott et al. [260] predicted the stability of buckled silicene monolayers on Cl-passivated Si(1 1 1) and clean
CaF2(1 1 1) surfaces using DFT calculations. For both surfaces, the lattice mismatches are small (0.1% and 0.5%) and the weak
van der Waals (vdW) silicene–substrate interaction preserves well the intrinsic electronic properties of freestanding silicene
sheet (e.g., Dirac cone and high Fermi velocity).
Using first-principles calculations, Das and co-workers [258] explored the bonding, stability, and electronic structure of
silicene monolayer epitaxially grown on various semiconductor substrates with bare surfaces: AlAs(1 1 1), AlP(1 1 1), GaAs
(1 1 1), GaP(1 1 1), ZnS(1 1 1), and ZnSe(1 1 1). They found that the relative stability and other physical properties of the
silicene layer rely sensitively on whether the interacting top layer of the substrate is metal or non-metal terminated.
Depending on the surface work function (WF), the silicene undergoes n-type (p-type) doping on metal (nonmetal) termi-
nated surface of the semiconductor substrates. The binding energy between silicene layer and metal-terminated surfaces
of these substrates falls in the range of 0.56 ± 0.12 eV/atom, which is comparable to that on Ag(1 1 1) surface (0.52 eV/atom);
while the binding energy lies between 0.37 and 0.92 eV/atom for the nonmetal-terminated surfaces.
Graphene bilayer was also proposed as a scaffold to synthesize silicene with electronic properties decoupled from the
substrate [259]. First-principles calculations demonstrated that the electronic and atomic structure of silicene intercalated
by graphene layers resemble those of standalone buckled silicene due to weak vdW interaction between silicene and
graphene layers. AIMD simulations further revealed that the in-plane hexagonal lattice structure of confined silicene is
preserved at room temperature and the silicene layer is stable even beyond 1000 K.

3.3. Silicene nanoribbons

Tailoring graphene into nanoribbons makes it possible to create an energy bandgap, owing to the quantum confinement
within the finite ribbon width [261,262]. One may expect similar effect if silicene is patterned into nanoribbons. In addition,
the buckled structure of Si bonds makes it possible to tune the electronic property of silicene by an external field. Combining
the size effect and external field, silicene nanoribbons (SiNRs) have been theoretically predicted to possess many extraordi-
nary properties, such as size-dependent energy gap [8,263], magnetic ordering [8] and unusual transport properties
[264,265].
Interestingly, in contrast to graphene where nanoribbons are difficult to obtain, silicene nanoribbons (SiNRs) have been
found to form spontaneously, and perfectly ordered on Ag(0 0 1) [14], Ag(1 1 0) [15–17,46,266–271] and Au(1 1 0) surfaces
[18]. The experimental reports of SiNRs on Ag(0 0 1) and Ag(1 1 0) surfaces were even prior to the finding of silicene sheet
on Ag(1 1 1) substrate; but whether these SiNRs are pure and pristine silicene ribbon structures remains an open question.
The SiNRs show metallic behavior, and quantum well states induced by confinement in the ribbon width direction were
indeed observed [18,272]. In an earlier ARPES study, it was also reported that SiNRs on Ag(1 1 0) exhibit a Dirac cone at the X
point of the Brillouin zone [17], suggesting that they could possibly be graphene-like nanoribbons. In this section, we will
mainly focus on the experimental results of SiNRs on Ag(1 1 0), including their structure models, electronic states, as well
as the remaining debates in these systems. The theoretical aspects of silicene nanoribbons will be summarized in Section 4.5.
Other details about SiNRs can be found in a few previous review articles [46,267,273].
SiNRs can form on Ag(1 1 0) substrate at a wide range of substrate temperature, from room temperature to about 500 K.
There are two types of SiNRs. At room temperature, narrow ribbons with two rows of protrusions in STM images are formed.
These nanoribbons exhibit identical width of about 1.0 nm and length from a few nanometers to a few tens of nanometers
[268,269,271]. When the substrate temperature is raised, the second type of nanoribbon with four rows of protrusions in the
STM image forms and dominates the surface as the substrate temperature reaches above 480 K. The width of these wider
ribbons is 2.0 nm and their length can extend up to hundreds of nanometers, limited mainly by the size of the Ag(1 1 0)
terraces. Both types of SiNRs are running along the ½1  1 0 direction of the Ag(1 1 0) surface. When properly annealed, the
SiNRs can also pack tightly in the [1 0 0] direction and fully cover the surface to form a monolayer SiNR grating. After further
increasing the Si coverage beyond monolayer coverage, multilayer structures will form, which preserve the high anisotropic
shape of the nanoribbons [268,274].
J. Zhao et al. / Progress in Materials Science 83 (2016) 24–151 71

Fig. 50. (a) 100  100 nm2 STM topographic image (V = 1.0 V) of 0.3 ML silicon deposited on Ag(1 1 0) at room temperature. (b) High resolution STM
topographic image (15  15 nm2, V = 1.1 V) of the structure of the 1.0 nm wide SiNRs. (c) 140  140 nm2 STM topographic image (V = 1.6 V) of 0.4 ML silicon
deposited on Ag(1 1 0) at 440 K. (d) High resolution STM topographic image (15  15 nm2, V = 1.0 V) of the atomic structure of the 2.0 nm wide SiNRs. (e)
High resolution STM image showing the atomic structure of Ag(1 1 0) and the SiNRs simultaneously. (f) Line profile across the white line in (d) showing the
width of the nanoribbon. Reproduced with permission from Ref. [271]. Copyright 2016, Elsevier B.V.

As shown in Fig. 50, the high resolution STM images of the 2 nm SiNRs consist four rows of protrusions. The periodicity
 0 direction. The protrusions
along the nanoribbon is exactly 2aAg, where aAg is the lattice constant of Ag(1 1 0) along the ½1 1
along two edges of the ribbon are shifted half of a period along the ribbon direction with respect to each other. For the 1 nm
ribbons that have only two rows of protrusions, the two rows appear similar as the two edges of the 2 nm ribbons, with the
72 J. Zhao et al. / Progress in Materials Science 83 (2016) 24–151

same 2 periodicity along the ribbon and shifted half of a period with respect to each other. This is consistent with the LEED
observations where the formation of SiNRs, a sharp 2 feature immediately appear in the ½1  1 0 direction [269]. On the other
hand, packing of the SiNRs in the [1 0 0] direction is not so well ordered. In the optimum annealing condition, the 1 nm and
2 nm SiNRs show 3 [269] and 5 [17,46,268,269,271] periodicities, respectively, along the [1 0 0] direction.
Note that there are some controversies in the early literature on SiNRs. For examples, Kara et al. [267] and Aufray et al.
[15] reported only one type of SiNR (2.0 nm), which they claimed to form already at the room temperature. In the STM
images, these ribbons exhibit three rows of protrusions, while in the high-resolution STM images they could directly reveal
four rows of honeycombs reflecting graphene-like ribbons. These results were proven incorrect by following experiments by
other groups [268,269,271]. The reason they only observed the 2.0 nm SiNRs could be due to the heating of substrate by the
irradiation from the high temperature silicon source. The abnormal STM images they observed could be due to both tip
artifacts and un-adjusted thermal drift.
The key issue in this kind of systems is to understand the atomic structure and composition of the two types of SiNRs.
There have been several theoretical proposals for the structural model of 2 nm SiNRs on Ag(1 1 0). Kara et al. proposed a
zigzag-edge model for the 2 nm wide SiNRs [15,267,275]. However, their model has a rectangular unit cell, and the period
along the nanoribbon is one lattice constant of silicene-1  1, i.e. 0.38 nm. These features are obviously in contradiction with
the STM image of the 2 nm SiNRs reported by other group [271], as shown in Fig. 50e. Another structural model proposed by
Lian et al. is the armchair silicene nanoribbon [276]. This model produces a period of 5.87 Å along the SiNRs, which agrees
with the experimental value. However, the LDOS simulations of this model also showed a rectangular symmetry, in contrast
to the experiments. These structures are based on a pure un-reconstructed silicene ribbon overlapping on the Ag(1 1 0)
substrate along the [1 1 0] directions. Recently, Feng et al. [271] proposed edge-reconstructed models for both the 1 nm
and 2 nm SiNRs, as depicted in Fig. 51, which can nicely explain the observed features in STM images. In their model, the
p p
center part of the nanoribbon is honeycomb, buckled silicene structure similar to the 3  3 structure of monolayer
silicene on Ag(1 1 1). The edges are reconstructed armchair type, the formation of which is probably due to the tendency
for Si atoms to form sp3 hybridization instead of sp2 hybridization.
Unlike some incommensurate structure of monolayer silicene on Ag(1 1 1) substrate, silicon nanoribbons form commen-
surate structure on Ag(1 1 0), as evidenced by both the STM images and the 5  2 LEED patterns (in the case of 2 nm SiNR). As
a freestanding silicene sheet has a smaller lattice constant (3.86 Å) than the column–column distance (4.08 Å) on the Ag
(1 1 0) surface along the [0 0 1] direction, the tensile stress caused by the mismatch between the lattices of silicene and Ag
substrate increases with the increasing width of the SiNRs along Ag[0 0 1] direction. This may explain why the experimen-
tally observed width of silicene ribbon is limited to narrower than 2.0 nm on Ag(1 1 0) surface.
Along with their discovery of SiNRs on Ag(1 1 0), De Padova et al. also reported the ARPES measurement of a surface fully
covered by 2.0 nm SiNR grating (Fig. 52) [17]. Two major features are revealed in the spectrum. First, there are several promi-
nent states (S1–S4) associated with the SiNR surface. These states disperse along the (C ! X) direction (the direction parallel
to the SiNRs), while they are dispersionless in the direction perpendicular to the SiNRs. These states can be separated well
from the strong sp band of the Ag substrate, and can be well explained by the quantum well states (QWSs) due to the lateral
confinement within the narrow ribbon width. Note that just beside the Ag sp band there are two cone-like Si bands near the
X point of the Ag surface BZ, separated by a gap of 0.6 eV. An attractive explanation of these bands is that they correspond to
the Dirac state originating from the p and p⁄ states, whereas the gap opening may be due to either interaction with the
substrate or confinement effect. The observed QWSs have been also confirmed by STS measurements, which reveals both
the spatial and energy distributions of the states [271,277]. The energy spectrum from the STS measurement agrees with
the S2–S4 states in the ARPES result in Fig. 52.
Fig. 53a–d shows the QWSs measured by STS on a separated 2 nm SiNR. At low bias, the LDOS is mainly distributed
around two edges of the SiNR, with the center being depressed (Fig. 53a). With increasing bias voltage, the LDOS at the edges
disappear. Meanwhile, one, two and three bright strings appear at the center of the nanoribbon successively (Fig. 53b–d). The
line profiles across the nanoribbon for images measured from 0.5 V to 4.0 V show the evolution of peaks from one, two to
three with increasing the bias voltage, as indicated by the black dotted lines in Fig. 53e. The oscillating patterns inside the
SiNRs can be assigned to quantum well states, as explained by an 1D ‘‘particle-in-a-box” model [271].
Recently, there have been some debates on the growth of Si on Ag. Several papers reported that the morphology of the Ag
(1 1 0) substrate is substantially modified during the growth of Si [269,278], suggesting possible alloying of Si with Ag. The
condition is similar to the case of silicene sheet on Ag(1 1 1) surface, where the Ag(1 1 1) substrate has also been modified
during the growth of Si [188]. However, as we will discuss in Section 4.1, both theoretical calculations and extensive exper-
iments seem to prove the validity of a pure silicene model consisting of 3  3 buckled Si atoms on Ag(1 1 1), without alloying
with the Ag(1 1 1) substrate. It is also worth to note that on both the Ag(1 1 0) and Ag(1 1 1) substrates, silicon atoms desorb
completely after annealing to about 700 K. Therefore, we suggest that the most probable structure of SiNRs on Ag(1 1 0)
consists of only Si atoms. Further evidences are desired to confirm the composition of the as-prepared SiNRs on Ag(1 1 0).
The second issue is the atomic structure of the SiNRs. Although it is attractive to interpret the SiNR as a cut of a pure
silicene sheet, the existing structural models based on pure, unreconstructed silicene ribbon cannot explain the observed
STM images. In addition, a recent Raman study also indicated that the excitation spectrum is not consistent with a honey-
comb silicene structure [277]. Silicene nanoribbons with reconstructed edges could be a possible solution to this puzzling
issue, while Ag–Si alloying also remains another possibility. Combining STM imaging and DFT calculations, Hogan and
J. Zhao et al. / Progress in Materials Science 83 (2016) 24–151 73

Fig. 51. (a and b) Top and side view of the relaxed structural model of 1 nm wide SiNRs on top of Ag(1 1 0). (c and d) Top and side view of the relaxed
structural model of 2 nm wide SiNRs on top of Ag(1 1 0). Light blue balls: topmost Ag atoms; dark blue balls: underlying Ag atoms; red balls: upper buckled
silicon atoms that can be probed by STM; yellow balls: other silicon atoms. (e–l) Simulated STM images compared with the experimental STM images at
different bias voltages. Bias voltage: (e) 1.1 V; (g) 1.0 V; (i) 1.5 V; (k) 1.0 V; (f and j) 0–1.5 V; (h and l) 1.5 to 0 V; (m and n) Calculated partial density of
states of the relaxed SiNRs without Ag(1 1 0) for the 1 nm and 2 nm SiNRs, respectively. Reproduced with permission from Ref. [271]. Copyright 2016,
Elsevier B.V.

Ó American Institute of Physics 2016

Fig. 52. (a) Energy distribution curves for bare Ag(1 1 0) and for the array of SiNRs; (b) band dispersion for the array of SiNRs vs. kx (along the SiNRs) at
ky = 0.7 Å1; (c) and vs. ky (perpendicular to the SiNRs) at kx = 0.35 Å1. Reproduced with permission from Ref. [17].

co-workers recently proposed two reconstruction models for Si nanoribbons on Ag(1 1 0) [279]. For coverage less than 1 ML,
the main structural motif is a zigzag chain of Si adatoms backbonded to Si dimers lying within the Ag missing rows. Multiple
lateral repetitions of this motif explain the existence of single, double, and triple nanoribbons.

3.4. Multilayer silicene

A stacking of graphene layers would result in few layer graphene, and eventually graphite with the increasing number of
layers. Owing to the nature of sp2 hybridization of carbon, the interlayer interaction between graphene layers is weak vdW
interaction, and thus the 2D nature of graphene is mostly preserved even in multilayer graphene systems. Following the
successful preparation of monolayer silicene on different substrates, it is natural to investigate the structure and properties
of multilayer silicene. There have been many theoretical studies on the bilayer [35,280–286] and multilayer [217,287–289]
silicene sheets using first-principles methods, mainly focusing on the stacking modes, interlayer interactions, and electronic
properties.
While monolayer silicene have been grown also on other substrates such as Ir(1 1 1), ZrB2 and MoS2, multilayer silicene
can be fabricated only on Ag(1 1 1) substrate so far [20,25,26,179,180,182,213–215,217,290,291]. Feng et al. [20] first
p p
reported the existence of bilayer and multilayer silicene films, whose surfaces exhibit a 3  3 honeycomb superstructure
with respect to the Si(1 1 1) lattice, independent of the film thickness. The multilayer silicene film can be grown on an initial
74 J. Zhao et al. / Progress in Materials Science 83 (2016) 24–151

Fig. 53. dI/dV maps (20  20 nm2) on an isolated 2.0 nm nanoribbon at different bias voltages. (a) 0.8 V. (b) 1.2 V. (c) 2.8 V. (d) 3.8 V. (e) Line profiles
perpendicular to the SiNR. The vertical coordinates are offset from one another for clarity. The red curves (also indicated by black arrows on the right)
represent pure eigenstates of the quantum well; the black dotted lines are guides to the eyes for the peaks. (f) Energy–momentum dispersion of the
electronic states for the 2.0 nm quantum well with each point obtained from the red curve in (e). The red solid line are the best parabolic fit to the data,
which yields the onset energy E0 = 1.36 eV and effective mass m⁄ = 0.5me. Reproduced with permission from Ref. [271]. Copyright 2016, Elsevier B.V.
J. Zhao et al. / Progress in Materials Science 83 (2016) 24–151 75

Fig. 54. (a) STM image (Vtip = 0.4 V, I = 100 pA, 45  45 nm2) obtained on top of Si film of 20 ML, containing an island with both armchair and zigzag step
p p
edges. Inset: the atomic model of 3  3 superstructure of Si. The black dash hexagon shows a honeycomb unit cell. The first Brillouin zone of the Si(1 1 1)
1  1 lattice is also shown. (b) dI/dV map (Vtip = 0.4 V, I = 200 pA, 45  45 nm2) of the same area as (a) showing obvious standing wave. (c and d) Line
profiles of LDOS along the x axis for armchair and zigzag edges labeled in (b) at various energies, respectively. The back dash lines are experimental values,
and the red lines are the lines fitting to the data. (e) Energy–momentum dispersions (E–k) determined from wave length of standing waves from armchair
and zigzag edges in (c and d), respectively. The Energy–momentum dispersion obtained from Ag(1 1 1) surface state is also shown for comparison. (f) E–k
curves (armchair edge) obtained on surface of Si films with different thickness on Ag(1 1 1) surface [213].

p p
monolayer silicene surface as a buffer layer, and the successive top layers all display 3  3 superstructure [214,215]. The
p p
appearance of 3  3 superstructure on the top surface of multilayer silicene has been discussed by theoretical calcula-
tions [231,292]. For example, Guo et al. [231] calculated the structure stability of bilayer silicene and found three symmetric
configurations with similar binding energy. The energy barrier for transformation among the three structures is small, and
thus flip-flop motions may occur above a critical temperature, while below this temperature the system may exhibit
p p p p
symmetric 3  3 ground state. It should be mentioned that the 3  3 honeycomb superstructure was at the beginning
reported as a monolayer silicene structure [175], and explained by a flip-flop motion between bistable states at higher
temperature [218]. However, based on the growth behavior of multilayer silicene on monolayer silicene, it is now commonly
believed that these observations should be ascribed to bilayer silicene.
76 J. Zhao et al. / Progress in Materials Science 83 (2016) 24–151

Ó IOP Publishing 2016

p p
Fig. 55. Dirac cones for multilayer ( 3  3) R30° silicene islands on top the first (3  3) wetting layer. (a) Left: scheme of the Brillouin zones of (1  1) Ag
p p
(1 1 1) (blue hexagon), (1  1) standalone silicene (black hexagons), ( 3  3) R30° silicene (green hexagons) and (3  3) silicene (red hexagons). Right:
scheme of a Dirac cone, with C- and V-shape linear p and p silicene bands, where the green lines represent the dispersion along the C0 ! K silicene direction.

p p
(b and c) Top: band dispersions around, measured along the C0 ! K silicene direction, collected at 1.5 ML, (3  3) plus ( 3  3) symmetries in LEED patterns,
p p
and 3 MLs (dominant 3  3 symmetry). Bottom: their waterfall line profiles. For clarity, we traced the dashed green lines in (b) (not in (c)), which are the
C- and V-shape linear p and p⁄ silicene bands, and the dashed white lines B, which are the structures related to the (3  3) symmetry. Reproduced with
permission from Ref. [26].

p p
The electronic structures of bilayer silicene with 3  3 honeycomb superstructure were first examined by STS
[175,238]. Pronounced quasiparticle interference patterns can be observed in the dI/dV maps at different tunneling bias
voltage. Linear energy–momentum dispersion relation can be extracted from the QPI patterns, which supports the existence
of Dirac fermions. Recently, Chen et al. [213] investigated the surface of multilayer silicene films with thickness up to 40
monolayers, and observed similar QPI patterns from all films with different thicknesses. This indicates that the metallic
p p
surface state on our Si(1 1 1) surface is delocalized and originates from the 3  3 superstructure on the surface rather than
the Ag(1 1 1) substrate. Fig. 54 shows the curves along C–K for films with thickness of 4, 8, 12 and 31 ML. All of them exhibit
linear energy–momentum dispersion with the same slopes, but are right shifted with increasing thickness. A possibility to
J. Zhao et al. / Progress in Materials Science 83 (2016) 24–151 77

Fig. 56. (a and b) The STM images of same area (Vtip = 1.0 V, I = 100 pA, 40  40 nm2) taken before and after applying a series of bias pulses (5 V, 50 ms) at
the position marked by the black dot in (a), respectively. (c) The line profiles along the blue lines in (a and b) respectively. (d) 3D version of (b) which
indicate the atomic structures of top layer [295].

account for the thickness-dependent energy shift is based on a double barrier tunneling junction model. The STM tunneling
junction involves two parts: one junction is between the STM tip and the Si film surface, and the other between the Si film
and the Ag(1 1 1) substrate. The bias voltage applied between sample and tip will be divided onto two junctions. The mag-
nitude of voltage drop in each junction is proportional to the resistance of junction. When the thickness of Si film increases,
the resistance of the junction between Si and Ag also increases. Thus the voltage drop between the STM tip and the Si film
surface decreases. As a result, the measured electronic band will be lowered. However, a quantitative explanation still needs
more theoretical efforts.
Using ARPES technique, De Padova et al. have investigated the electronic structures of multilayer silicene films with
different thicknesses [25,26]. Different from monolayer silicene where its electronic structure may be significantly affected
by the substrate, they have observed clear Dirac cone structure at the C point of the surface Brillouin zone in silicene films
with different thicknesses. As depicted in Fig. 55, the quasilinear p⁄ and p bands meet at 0.25 eV below the Fermi level with
an approximate Fermi velocity of 0.3  106 m/s, providing clear evidence of the presence of gapless Dirac fermions. Such
bands do not coincide with any of the bulk bands of Ag(1 1 1) substrate and should be ascribed to the silicene film only. The
thickness independent character further indicates that the Dirac cone may stem mainly from the top surface layer. A recent
p p p p
ARPES study of 3  3 silicene layer on top of the 13  13/4  4 buffer layer grown on Ag(1 1 1) substrate also revealed
a Dirac cone structure with the Dirac point 0.33 eV below the Fermi level [182].
On the contrary, combining soft X-ray emission and absorption spectroscopies at the Si L2,3-edge and DFT calculations,
Johnson and co-workers [290] demonstrated that epitaxial silicene bilayers on the Ag(1 1 1) surface are metallic and strongly
interact with the underlying substrate, without showing a Dirac cone at the Fermi level. Unlike graphite where the carbon
atoms are entirely sp2 hybridized in a planar structure, in silicene the silicon atoms are partially sp2/sp3 hybridized in a buck-
led structure. It is well accepted that sp3 hybridization favors a diamond-like structure rather than a 2D layered structure.
Indeed, STM investigations indicated that the interlayer spacing in multilayer silicene is about 3.14 Å, very close to the inter-
layer spacing of bulk Si(1 1 1) planes in diamond structure of silicon solid. One the other hand, high resolution STM images
reveal that most films are of ABC stacking, instead of the AA stacking in graphite [213]. These results are evidences of strong
interactions between silicene layers, although whether it is a purely diamond structure is still under debate. For example,
78 J. Zhao et al. / Progress in Materials Science 83 (2016) 24–151

Raman studies have indicated that the sp3 nature of silicon bonds is largely relaxed in multilayer silicene [182]. Moreover, all
theoretical calculations of multilayers silicene found strong covalent bonding between the silicene layers with the absence of
the Dirac cone, as opposed to weak vdW interaction between the graphene layers [35,280–288].
p p
Another debating issue on multilayer silicene is whether the 3  3 structure on the top surface of multilayer silicene is
p p
pristine silicon structure, or just a well-known 3  3 phase formed by a monolayer Ag atoms on a bulk Si(1 1 1) substrate,
which is conventionally observed on a single crystalline Si(1 1 1) surface with 1 ML Ag adatoms. In an experiment by Boren-
sztein et al. [293] using LEED, Auger spectroscopy, and optical reflectance methods, it was argued that Si multilayers grown
p p
on Ag(1 1 1) at high temperatures are not multilayer silicene, but cubic diamond-like silicon. The 3  3 reconstruction is
induced by about 0.5 ML of remanet Ag atoms at the surface. Recently, Shirai et al. [294] performed LEED IV measurement
p
and derived that the structure of the surface is identical to the well-known inequivalent triangle (IET) model of Si(1 1 1)-
p
3  3-Ag. Combing with the fact that the stacking structure of multilayer silicene is quite close to the bulk Si(1 1 1), it was
p p
suggested that the 3  3 structure is formed due to the segregation of Ag atom to the Si(1 1 1) film surface. If this is the
case, the so-called ‘‘multilayer silicene” would be nothing more than a Si(1 1 1) film with Ag-terminated surface. In STM
p p p p
experiments, the 3  3 phase and Si(1 1 1) 3  3-Ag surface actually exhibit very similar feature.
However, this argument contradicts at least with several experimental facts. Firstly it cannot explain the Dirac cone struc-
p
ture observed by APRES [25,26,182], which is definitely different from the ARPES spectrum of the well-known Si(1 1 1)-
p
3  3-Ag. Secondly, Chen et al. [295] recently showed the possibility to peel the silicene layers from multilayer silicene.
p p
A series of repeated bias pulses were applied to the tip on the surface of 3  3 phase, and the top layer beneath the
tip was damaged and removed, leaving a pit inside which the underneath layer was exposed, as shown in Fig. 56. The line
profiles indicate that the area inside the pit is about 0.62 nm lower than the original surface, corresponding to the removal of
two layers from the surface. The 3D derivative STM image in Fig. 56d shows the atomic structures of the exposed layer is
p p p p
identical to the top layer and exhibits a 3  3 reconstruction. Since the Si(1 1 1) 3  3-Ag surface consists of only
one layer of Ag atoms on top of the pure Si(1 1 1) substrate, one should not observe the reconstruction on the underneath
p p p p
layer when the top layer is removed. Therefore, the 3  3 reconstruction is not the Ag–Si(1 1 1)-( 3  3)R30°, but is
p p p p
an intrinsic structure of pure silicon. Additionally, there are other differences between 3  3 and Si(1 1 1)( 3  3)-
Ag. For example, although both surfaces exhibit a structural phase transition at low temperature, the transition temperature
p p p p
is distinctly different: 30–40 K for the 3  3 phase [218], and above 150 K for Si(1 1 1)( 3  3)-Ag [296].
A possible explanation to these contradicting experimental results is that there exists a narrow temperature window for
p p
multilayer silicene growth, where Ag atoms do not segregate to the surface. Since a Si(1 1 1)- 3  3-Ag reconstruction
forms at a relatively high temperature, it is suggested that one should keep the substrate temperature slightly below the
p p
formation temperature of Si(1 1 1)- 3  3-Ag, while still maintaining the sufficient diffusion of silicon atoms on the
surface to form a flat silicon film. We suggest that in order to obtain a multilayer silicene film, one should first prepare a
monolayer silicene film as the wetting layer, and then the multilayer film can be grown at a temperature of about
p p
200–250 °C, which is below the formation temperature of Si(1 1 1)- 3  3-Ag (above 700 K).
Recent studies revealed strong influence of the substrate on the electronic structure of monolayer silicene, and that the
Dirac state could no longer exist in monolayer silicene on metal substrate. This is a serious challenge to any further research
and application of silicene. Alternatively, multilayer silicene could provide a more favorable system because the Dirac-like
p p
electronic properties stems from the 3  3-reconstructed surface and is more robust, and the metal substrate is isolated
p p
by the underlying layers. Moreover, the 3  3 surface of multilayer silicene was found to be surprisingly stable to sustain
a 24 h exposure to air [291]. Therefore, it is expected that multilayer silicene may play a role in fabrication of silicene devices
in the future, taking the advantage of mature silicon micro- and nano-technologies.
Very recently, Yaokawa et al. [297] reported fabrication of a novel form of bilayer silicene (denoted as w-BLSi) from
calcium-intercalated monolayer silicene (CaSi2). This w-BLSi structure has not been theoretically considered in early
calculations of bilayer silicene [225,231,279,285,291]. In particular, the number of unsaturated silicon bonds in the
w-BLSi is reduced compared with monolayer silicene, and an indirect band gap of 1.08 eV is opened. This progress definitely
opens a new doorway towards synthesis of bilayer and maybe even multilayer silicene in addition to the conventional MBE
synthesis on Ag(1 1 1) substrate.

4. Modification of silicene

4.1. Hydrogenation

The zero-gap character of silicene hinders its applications in nanoelectronic and optoelectronic devices. It is thus
desirable to open a finite band gap in silicene. Hydrogenation can convert sp2 hybridized silicon into sp3 and provide a simple
way to fulfill the goal of gap opening [37,298]. Fully hydrogenated silicene is named as silicane. As presented in Fig. 57, two
different configurations of silicane, i.e. chair and boat, are found to be stable, similar to that of graphane [299,300]. In the
chair configuration, neighboring hydrogen atoms sit alternatively above and below the basal plane, while the boat one
has paired hydrogen atoms alternating. According to the DFT computations, chair configuration is more stable than boat
configuration by 25 meV/atom [37] or 30 meV/atom [301]. For graphane, chair configuration is also more stable than boat
configuration by 60 meV/atom [299] or 51 meV/atom [300]. After full hydrogenation, the bucking height of silicene sheet
J. Zhao et al. / Progress in Materials Science 83 (2016) 24–151 79

increases from 0.44 Å to 0.72 Å [37], indicating an enhanced sp3 hybridization. The formation energy of silicane with chair
configuration is calculated to be only 0.032 eV per atom with respect to bulk silicon solid and H2 gas, which is much smaller
than that of monolayer silicene (0.779 eV per atom) [35]. For comparison, the formation energy of graphane with chair
configuration is about 0.15 eV per atom. Thus hydrogenation can substantially stabilize these elemental monolayer sheets.
The dynamical stabilities of the chair and boat configurations of silicane sheets have been confirmed by phonon dispersion
curves from first-principles calculations [298,302–304]. From DFT calculation, the energy barrier for dissociative adsorption
of a H2 molecule on silicene sheet is 1.75 eV, which can be reduced to 0.38 eV by applying an electric field of 0.05 a.u. [305] It
was also found that the adsorbed H atoms on silicene basal plane tend to group into clusters stabilizing each other in a 2D
nucleation and growth mechanism [306].
To date, the electronic band structures of silicane have been extensively computed using DFT with different treatments of
exchange–correlation interactions [37,298,301,304,307–309], various TB models [302,310,311], as well as many-body
Green’s function approach within the GW approximation. The theoretical values of band gaps are summarized in Table 4
and the representative band structures from LDA calculations are shown in Fig. 58. Upon full hydrogenation, an energy
gap opens in silicene due to the sp3 hybridization of the Si atoms. Silicane with chair configuration is an indirect semicon-
ductor (Fig. 58a), while the boat one is a direct semiconductor (Fig. 58b). The band gaps of chair-like (boat-like) silicane
predicted from the more accurate HSE and GW methods are in the range of 3.5–4.0 (2.4–3.3) eV. By contrast, both the chair
and boat configurations of graphene are direct semiconductors with larger band gaps of 5.4 eV and 5.1 eV from GW
calculations, respectively [300]. For the chair-like silicane, the effective masses for electron and hole carriers are predicted
to be 0.04m0 and 0.076m0 (m0 is the electron mass), respectively [311].
The fundamental mechanical properties of silicane have been computed by Peng and De using first-principles method
[303]. Compared to silicene, silicane has relatively low in-plane stiffness (53.8 N/m) and low Poisson ratio (0.24), indicating
that full hydrogenation of the silicene moderately reduces its intrinsic strength and Poisson ratio. Due to the increased buck-
ling roughness of 59% by hydrogenation, the ultimate strains increase by 29%, 33%, and 24% under armchair, zigzag, and biax-
ial deformations, respectively. In contrast, hydrogenation has little effect on the ultimate tensile strengths of silicene sheet.
Apart from silicane with full hydrogenation on two sides, one-side semihydrogenated silicene have also been investigated
by first-principles calculations [301,307,313,314]. No imaginary phonon branch occurs in the phonon dispersions curve,
confirming its dynamical stability. As displayed in Fig. 59, there are three possible configurations of half-hydrogenated
silicene: (a) zigzag conformer with the H atoms absorbed on silicon zigzag chains alternately, (b) boat conformer with
the H atoms distributed in pairs, (c) chair conformer with the H atoms sitting alternately on only one side of the sheet.
According to the DFT calculations by Zhang et al. [301], the zigzag conformer is more stable than the boat and chair conform-
ers by 33 and 180 meV/atom, respectively.
The computed band gaps of half-hydrogenated silicene with various configurations are summarized in Table 4. Both
chair-like and boat-like semihydrogenated silicene sheets are semiconductors with direct electronic gaps
[301,307,313,314]. The band gaps from HSE06 calculations of 1.79 eV (chair) and 1.14 eV (boat) [301], are only about half
of the fully hydrogenated silicene with the same configuration. More interestingly, ferromagnetism is found in the
chair-like silicene with one-side hydrogenation, where every unsaturated Si atom carries a magnetic moment of 1 lB. On
the other hand, the most stable zigzag configuration is metallic and non-magnetic. The different magnetic orderings in

Fig. 57. Top view (upper) and side view (lower) of silicane in (a) chair and (b) boat configurations. The yellow balls represent Si atoms and the white balls
represent H atoms.
80 J. Zhao et al. / Progress in Materials Science 83 (2016) 24–151

Table 4
Band gaps (eV) of two-side and one-side hydrogenated silicene with different configurations (shown in Figs. 57 and 59) from the theoretical calculations using
different methods (GGA denotes PW91 or PBE).

Method Two-side hydrogenation One-side hydrogenation


Chair Boat Chair Boat Zigzag
TB 2.2 [310], 2.31 [311] 1.93 [311]
LDA 2.0 [37,298] 1.6 [298]
GGA 2.11 [312], 2.19 [304], 2.36 [301], 2.37 [309] 1.6 [301] 0.84 [301], 0.95 [307], 0.94 [313] 0.53 [301] 0 [301]
HSE06 4.0 [298], 3.51 [301] 3.3 [298], 2.41 [301] 1.79 [301], 1.74 [313] 1.14 [301] 0 [301]
GW 3.8 [298] 2.9 [298]

Ó American Institute of Physics 2016


Fig. 58. Energy band structures, calculated using LDA functional, for (a) chair silicane, (b) boat silicane. The reference (zero) energy level corresponds to the
top of the valence band. Reproduced with permission from Ref. [298].

Fig. 59. Top view (upper) and side view (lower) of the atomic structures for: (a) zigzag semihydrogenated silicene; (b) boat-like semihydrogenated silicene;
(c) chair-like semihydrogenated silicene. The yellow balls and white balls represent Si and H atoms, respectively. The black lines show the unit cell.

half-hydrogenated silicene conformers originate from coupling of the pz orbitals of the unsaturated Si atoms. In a word,
hydrogenation provides a novel way to tailor the electronic and magnetic properties of silicene and shows great potentials
for future nanoelectronics and spintronics. The spintronic devices based on half-hydrogenated silicene will be further
discussed in Section 5.3.2.
As a 2D semiconductor with substantial band gap, silicane offers a playground to further doping and functionalization. Pi
and co-workers [315] explored substitutional doping of boron or phosphorus atoms in the chair-like silicane with dopant
concentration between 1.4% and 12.5%. After B- or P-doping, silicane is transformed from an indirect-gap semiconductor into
a direct one, and the band gap reduces with increasing dopant concentration. Moreover, doping can reduce the effective
mass of holes and electrons and thus increase the carrier mobility in silicane. In addition, organic functionalization of silicane
with acetylene, ethylene, and styrene have been studied using DFT calculations and compared with the adsorption on hydro-
genated Si(1 1 1) surface [316]. The abstraction barriers of acetylene and ethylene in silicane are much lower than H–Si(1 1 1),
while no big difference is found in the case of styrene.
J. Zhao et al. / Progress in Materials Science 83 (2016) 24–151 81

Ó American Physical Society 2016


Fig. 60. (a) A large area STM image of hydrogenated a-4  4 silicene surface showing an ordered 4  4 structure. (b) Zoom-in STM image of hydrogenated
a-4  4 phase. The white rhombus marks an apparent unit cell of the structure. There are six bright protrusions in one HUC and one protrusion in the other
HUC. (c) STM image showing the comparison between the position of apparent UCs of clean and hydrogenated a-4  4. The red and white rhombuses
correspond to clean a-4  4 UC and the hydrogenated a-4  4 UC, respectively. A translation of the white UC (dot line) does not match with the red one.
(d) The clean a-4  4 surface is fully recovered after annealing the surface at 450 K. Reproduced with permission from Ref. [39].

Using first-principles calculations, Rupp et al. [317] have shown that the substitutional doping by N and B impurities is a
powerful technique to functionalize silicane and convert the pristine silicane into n-type and p-type semiconductor, respec-
tively. In addition, they also [318] demonstrated that chemical functionalization of various elements (N, P, S, Li, Na, K, Mg and
Ca) is able to enhance the photocatalytic water-splitting activity of silicane for the hydrogen evolution reaction (HER) and
oxygen evolution reaction (OER).
Beyond monolayer silicene, hydrogenation on the bilayer and few-layer silicene sheets have been theoretically explored
by several groups [316,319,320]. Using DFT calculations, Liu et al. [319] demonstrated that the electronic and optical
properties of hydrogenated few-layer silicene sheets strongly depend on the thickness and stacking modes. The formation
energy increases with decreasing number of layers due to the higher surface/volume ratio. Because of the well-known quan-
tum confinement effect, there is an increased trend of band gap with the reduction of layer thickness, leading to an obvious
blue shift of optical absorption edge in the hydrogenated few-layer silicene. For all three stacking modes considered, the
band gap from DFT calculations can be fit into an empirical function as number of layers (N) as:

Eg ðNÞ ¼ E0g þ F=N x ; ð22Þ

where Eg0 = 0.39 eV, F = 1.7, x = 1.1 for AA stacking; Eg0 = 0.59 eV, F = 1.6, x = 1.3 for AABB stacking; Eg0 = 0.61 eV, F = 1.8,
x = 1.1 for ABC stacking, respectively. With the same layer thickness, the ABC-stacking has the highest stability and the
largest band gap, which originates from the strong interactions among the silicene layers.
Combining first-principles calculations and the cluster-expansion approach, Huang et al. [316] demonstrated that
partially hydrogenated bilayer silicene sheets are a kind of promising optoelectronic material with a widely tunable band
gap. At low hydrogen concentrations, four ground states of single- and double-sided hydrogenated bilayer silicene, i.e.,
82 J. Zhao et al. / Progress in Materials Science 83 (2016) 24–151

Ó American Physical Society 2016


Fig. 61. (a and b) are the structural model and the simulated STM image of hydrogen-terminated a-4  4 and b-4  4 model, respectively. (c and d) are the
structural model and the simulated STM image of hydrogen-terminated a-4  4 and b-4  4 models, respectively. In (c) the white and red rhombuses
corresponding to the positions of apparent UCs of a-4  4 and b-4  4 phases, respectively, which are shifted relatively. Note that in (a and c), the lateral
position of Si atoms are fixed unchanged and only the buckling configuration has changed, resulting in the change of the position of apparent Ucs.
Reproduced with permission from Ref. [39].

Si28H6, Si24H6, Si16H4, and Si12H4, are dipole-allowed direct band-gap semiconductors with band gaps of 1–1.5 eV that are
suitable for solar photovoltaic applications. At high hydrogen concentrations, three hydrogenated bilayer silicenes with
well-ordered structures, i.e., Si8H4, Si16H12, and Si12H10, exhibit direct (or quasidirect) band gaps in the color range of red, green,
and blue, which might be used as light emitters for white LEDs. Recently, fully and partially hydrogenated monolayer silicene
sheets with different hydrogen concentrations have been theoretically explored by Lin and co-workers [321]. A delta-function-
like peak in the DOS was observed at Fermi level, whose intensity is weakened with the reduced H-concentration, suggesting an
effective way to identify the H-concentration in scanning tunneling spectroscopy experiments.
In an experiment by Wu’s group, monolayer silicene with Si(3  3)/Ag(4  4) superstructure on Ag(1 1 1) surface has been
hydrogenated [39]. The silicene of a-4  4 phase was hydrogenated by absorbing atomic hydrogen in UHV. Upon exposing to
a saturated hydrogen dose at room temperature, a perfectly ordered structure with the same 4  4 periodicity was observed,
as displayed in Fig. 60a for a large area image. The distance between the nearest bright spots is about 3.8 Å, corresponding to
the lattice constant of silicene-1  1. Intuitively, adsorption of H atoms on upper-buckled Si atoms only changes its degree of
buckling slightly. The resulting STM pattern after hydrogenation should resemble that of the clean a-4  4 surface of silicene.
However, high resolution image of the hydrogenated structure in Fig. 60b manifests two inequivalent HUCs: one with six
bright spots, another with only one bright spot in the middle. This is different from the a-4  4 silicene, but reminiscent
of the b-4  4 phase. This observation can be understood by the fact that the b-4  4 phase prefers to stay in the strained
area of pure silicene sheet. The adsorption of H atoms will increase the degree of buckling of Si atoms and thus increases
the strain, making the b phase more stable. As shown in Fig. 61, the perfect agreement between theoretical model and
experimental STM images also supports that both the a-4  4 and b-4  4 are pure silicon structures. Moreover, the fully
half-hydrogenated silicene sheet can be completely restored to its original state by annealing the sample to about 450 K.
The adsorption–desorption cycle can be repeated many times without degradation of the silicene film if the ultrahigh
vacuum system is clean enough.
J. Zhao et al. / Progress in Materials Science 83 (2016) 24–151 83

Ó Royal Society of Chemistry 2016


Fig. 62. Energy band structures for (a) F-silicene, (b) Cl-silicene, (c) Br-silicene and (d) I-silicene. The Fermi level is labeled with a dashed line. Reproduced
with permission from Ref. [323].

In a successive experiment of hydrogenating various silicene phases on a Ag(1 1 1) substrate, Wu’s group [322] found that
p p
only 2 3  2 3 phase can produce half-silicane with perfect 1  1 lattice, while the formation mechanism was discussed
with DFT calculations.
Recently, there have been some questions on whether the experimentally observed silicene superstructures on Ag(1 1 1) is
actually Ag–Si alloy phase. Prevot et al. [188] explored in-situ STM measurement during the growth of silicene films. They
found that the morphology of Ag(1 1 1) surface suffers a significant change during growth, which may correspond to strong
Ag mass transport. Hence, Ag atoms should be incorporated into the resulting silicene film to form Ag–Si alloy. However,
mass transport is not a direct proof of the incorporation of Ag in silicene film. A pure silicene model without Ag alloying
is still the most plausible model to explain the STM images of monolayer silicene so far. The atomic structure model of
the a-4  4 silicene is a very natural one and is energetically stable. Especially, the simulated STM image perfectly matches
the experimental one. In contrast, there is no alloy structure model that can explain the STM images of the 4  4 phase yet.
The fact that the hydrogen adsorption phenomena can be nicely explained by existing silicene model supports the pure
silicene model without Ag–Si alloying. Moreover, as hydrogen atoms do not react with silver atoms, each hydrogen adsorp-
tion site should correspond to one silicon atom, which is upper buckled. The hydrogenated silicene 4  4 unit cell consist of
two half unit cells, one with six H atoms in a local 1  1 pattern, and the other with one H atom. The local 1  1 structure is
so compact that it is hard to insert Ag atoms into it. Therefore, there is still no alloy model which can explain these existing
STM observations so far.

4.2. Halogenation

Besides hydrogenation, chemisorption of halogen elements is also able to open a band gap in silicene and has been inten-
sive studied. Similar to silicane, the chair configuration is more stable than the boat configuration for X-silicene (X = F, Cl, Br
and I). The band structures of the chair-like X-silicene systems are shown in Fig. 62. From sX-LDA calculations, all X-silicene
are direct semiconductors with band gaps of 1.469 eV, 1.979 eV, 1.950 eV and 1.194 eV for F, Cl, Br and I, respectively [323].
The variation trend of Eg can be attributed to the competition of Si–Si and Si–X bond strengths.
Among X-silicene, the F-silicene with chair configuration has been intensively studied by DFT calculations and many-
body Green’s function approach. The theoretical band gaps fall in a wide range, i.e., 0.70 eV (PBE) [304], 1.2 eV (M06L)
84 J. Zhao et al. / Progress in Materials Science 83 (2016) 24–151

[324], 1.469 eV (sX-LDA) [323], 1.6 eV (HSE) [324], 2.33 eV (GW0) [304]. Moreover, a new configuration, so called Z-line
(Fig. 63), is nearly energetically degenerate with the chair configuration for F-silicene [304]. In the Z-line conformer, F atoms
are adsorbed along zigzag lines alternatively on both sides of silicene. Thus, it has the buckling alternated by the zigzag lines,
losing the symmetry of pristine silicene. The Z-line configuration of F-silicene is also a direct semiconductor, with band
structure shown in Fig. 63. Its band gap (0.82 eV at PBE level or 2.64 eV at GW0 level) is slightly larger than that of the chair
conformer by 15%.
Apart from full fluorination, Wang et al. [325] explored half-fluorinated silicene sheets using spin-polarized DFT calcula-
tions and considered the zigzag, chair, and boat configurations. All three conformers are dynamically stable as evidenced by
the phonon dispersion curves. Among them, the zigzag configuration is the most stable, similar to the case of half-
hydrogenated silicene. From the HSE03 calculations, half-fluorinated silicene with zigzag and boat configurations are direct
semiconductors with band gaps of 0.411 and 0.824 eV, respectively. Meanwhile, the chair-like half-fluorinated silicene
shows antiferromagnetic behavior, with each unsaturated Si atom carrying a magnetic moment of 0.84 lB. Similar antifer-
romagnetism was found in half-brominated silicene, where the unsaturated Si atom has a local moment of 0.24 lB and the Br
atom also contributes a small spin moment of 0.11 lB [314]. Moreover, partial fluorination provides an effective way to
realize tunable band gap and doping in silicene-based materials. Combining first-principles calculations with the cluster
expansion approach, Huang and co-workers [326] demonstrated that the band gaps of homogenous SiFx alloys with
two-side functionalization could be continually modulated from 0 to 1.59 eV.
Using DFT calculations, Denis [324] considered symmetrically and asymmetrically (Janus) functionalized monolayer
silicene with covalent additions of H, CH3, OH and F. They found that the electronic structure of monolayer silicene can
be modulated from semiconducting (with band gap up to 3.2 eV) to metallic by controlling the type and concentration of
radicals. They further explored the stacking of two functionalized silicene sheets and found significant reduction of band
gap with regard to the monolayer systems. A similar computational study has been conducted by Shin and co-workers
[327], who predicted piezoelectricity in the hydrogen and fluorine co-decorated silicene sheets. Recently, Zhao’s group

Ó American Institute of Physics 2016

Fig. 63. (a) Atomic structure and (b) band structure of F-silicene in Z-line configuration. The yellow balls and light blue balls represent Si and F atoms,
respectively. Reproduced with permission from Ref. [304].
Ó IOP Publishing 2016

Fig. 64. Normal emission (# = 0°) Si 2p core levels for a grating of silicene NRs after increasing oxygen exposures. The LEED patterns (Ep = 57 eV) were
collected on the bare grating and after the oxygen exposition of 100 L. Reproduced with permission from Ref. [330].
J. Zhao et al. / Progress in Materials Science 83 (2016) 24–151 85

[328] further explored asymmetrically functionalized silicene (i.e., Janus silicene or X-silicene-Y) with a large variety of X/Y
combinations: H/F, H/COOH, H/CN, CH3/F, CH3/COOH, CH3/CN, CH3O/F, CH3O/COOH, CH3O/CN, NH2/F, NH2/COOH, NH2/CN,
OH/F, OH/COOH, OH/CN. They found that these Janus silicene sheets exhibit controllable band gap, depending on the type
and concentration of the X and Y groups.
Beyond atoms of halogen elements, adsorption of a magnetic superhalogen molecule MnCl3 on silicene has been proposed
by Wang’s group [329]. By definition, a superhalogen molecule possesses an electron affinity even larger than Cl. The isolated
MnCl3 molecule carries a magnetic moment of 4 lB, which is retained when it is deposited individually on the silicene sheet.
Interestingly, the silicene sheet adsorbed with two MnCl3 molecules on the same side is ferromagnetic with a total spin
moment of 8 lB, whereas that with two MnCl3 on the opposite side is antiferromagnetic. Silicene with adsorption of single
MnCl3 is half-metallic, whereas adsorption with two MnCl3 transforms silicene into semiconductor with indirect band gap
up to 0.63 eV. Certainly, adsorption of superhalogen molecule MnCl3 and perhaps the other suitable molecules has led a
variety of opportunities of tuning the electronic and magnetic properties of silicene.

4.3. Oxidization

Unlike graphene sheets that are hydrophobic and chemically inert, silicene sheets exhibit high chemical activities, which
potentially make their physical, chemical and electronic properties easily adjustable by chemical functionalization. Oxygen
adatoms and molecules can be used to probe and modulate local electronic states in silicene via adsorption process at atomic
level, due to its high chemical activity. Thus, oxidation of silicene into silicene oxide is considered as one of the major steps
towards functionalization of silicene. On the other hand, chemical stability of silicene exposed in air is a critical concern to its
device applications. To date, the oxidation of silicene sheets and nanoribbons have been extensively studied both experimen-
tally and theoretically.
In an early study, De Padova and co-workers [330] investigated the room temperature oxidation of silicene nanoribbons
grown on Ag(1 1 0) surface by high-resolution Si 2p core level photoemission spectroscopy and low-energy electron diffrac-
tion. As shown in Fig. 64, the Si spectra are practically unaffected up to an exposure of 100 Langmuir (L): the region between
1.0 and 4.5 eV below the Si 2p core level is completely free of oxidation states. The oxidation process starts only at very high
oxygen exposures up to 1000 L, typically 104 higher than that on Si(1 1 1) 7  7 surface, indicating the strong resistance of

Ó American Chemical Society 2016

p p p p
Fig. 65. STM images of oxidized silicene in (a) 4  4, (b) 2 3  2 3, and (c) 13  13 structures (scanning area 4  4 nm2, Vbias = 0.2 V, I = 4 nA). The
oxygen adatoms prefer to reside on top-layer Si atoms in the initial oxidation. (d) Line profiles of oxygen adatoms on silicene corresponding to the lines in
the STM images in (a, b, and c), respectively. (e–g) DFT simulations (top and side views) of atomic structures for oxygen adatoms on Ag(1 1 1) supported
p p p p
silicene monolayers in different superstructures: (e) 4  4, (f) 2 3  2 3, (g) 13  13. The black rhombuses in the top views represent the unit cell.
Red: oxygen; yellow: silicon; blue: silver. Reproduced with permission from Ref. [40].
86 J. Zhao et al. / Progress in Materials Science 83 (2016) 24–151

silicene nanoribbons towards oxidation. The oxidation uptake can be enhanced by about two orders of magnitude by inten-
tionally created defects via Ar+ sputtering.
p p
Later, oxidation of epitaxial silicene sheets (with domains of 4  4 and 13  13 superstructures) grown on an Ag(1 1 1)
substrate was investigated by monitoring the Si 2p XPS photoemission line [331]. Silicene sheets were found to be chemi-
cally stable with O2 dosing (up to 1000 L) in an ultrahigh vacuum environment but undergo oxidation upon exposure to air.
The latter effect would be a serious drawback for device applications. The oxidation can be prevented by encapsulations of
the silicene layer with an ultrathin Al or Al2O3 of a few nm thickness. In contrast, Friedlein et al. [332] deposited Al atoms
onto epitaxial silicene on thin ZrB2(0 0 0 1) films grown on Si(1 1 1) and found that the aluminum oxide overlayers formed on
silicene do not prevent but actually promote the oxidation of silicene. Under oxygen exposure up to a dose of 4500 L, the bare
silicene is hardly affected, whereas Al-covered silicene becomes partially oxidized due to dissociative chemisorption of O2
molecules by Al atoms.
Combining measurements by Auger electron spectroscopy, X-ray diffraction and Raman spectroscopy, De Padova [291]
p p
demonstrated that thick epitaxial multilayer silicene films with a 3  3R(30°) surface structure (without any protective
capping) show only mild surface oxidation after 24 h exposure to ambient air. Comparison of AES spectra shows that even
after 24 h in air, the 43 MLs silicene film remains largely intact, being oxidized only at its top part. Moreover, the character-
istic G band peaked at 523.26 cm1 and the broad D band between 430 and 500 cm1 were observed in the ex-situ Raman
spectrum of multilayer silicene film in air, which locate at the same positions for the Al2O3 capped multilayer silicene film
and can be compared to the recent in-situ Raman feature of bilayer silicene on Ag(1 1 1) [182].
The stability of freestanding silicene sheets under oxygen environment has also been theoretically explored by Liu et al.
using first-principles calculations [333,334]. On pristine silicene, an O2 molecule can be easily adsorbed and then dissociate
into two O atoms without overcoming any energy barrier. The dissociation reaction is an exothermic process, and the
dissociated O atoms form strong Si–O bonds, lowering the energy by 4.046–5.355 eV per O2. For comparison, fully

Ó Macmillan Publishers Limited 2016

Fig. 66. Energy vs. k dispersion measured by ARPES for (a) clean Ag(1 1 1) surface, (b) 4  4 silicene grown on Ag(1 1 1), and (c) oxidized silicene on Ag(1 1 1),
respectively. SSS in (a and b) denotes the Shockley surface state. HSB in (b) denotes the hybrid surface band. (d) Schematic of BZ for 4  4 silicene grown on
Ag(1 1 1): blue and orange honeycomb structures correspond to Ag(1 1 1) and 4  4 silicene, respectively. Reproduced with permission from Ref.
[336].
J. Zhao et al. / Progress in Materials Science 83 (2016) 24–151 87

hydrogenated silicene is more stable upon oxidization than pristine silicene due to the existence of two energy barriers, i.e.,
0.268 eV between the initial and transition states and 0.249 eV from the transition and final states, respectively.
Recently, Morishita and Spencer [335] investigated the microscopic mechanism of the reaction of O2 with 4  4 silicene
on Ag(1 1 1) substrate using AIMD simulations. They found that there exist barrierless oxygen chemisorption pathways
around the outer Si atoms of the silicene overlayer, thus oxygen can easily react with a Si atom to form an Si–O–Si config-
uration. When the oxidation process involves multiple O2 molecules (corresponding to a high oxygen dose), a synergistic
effect between the molecular dissociation and subsequent structural rearrangements is found to accelerate the oxidation
process and to enhance self-organized formation of sp3-like tetrahedral configurations consisting of Si and O atoms. Their
theoretical results are helpful for explaining the previously reported inconsistent experimental observations of the oxidation
of silicene on Ag(1 1 1) substrate, where different flux (or pressure) and temperature of oxygen gas could induce distinct
oxidation processes.
Like its graphene counterpart [12], oxidization provides an effective way of tuning band gap of silicene in a wide range. It
is expected that local electronic structure can be tuned from zero-gap state to semiconducting state by changing the oxygen
level and the adsorption site of oxygen adatoms on silicene. Pi’s group [38] investigated the structures, thermodynamic
stability, and electronic properties of silicene oxides by incorporating atomic oxygen (O) and hydroxyl (OH) onto silicene
lattice in a variety of bonding configurations. They found that the oxidation of silicene in the atmosphere of oxygen gas most
likely leads to the formation of the single-side over-bridging configuration, while boat-like OH and umbrella-like OH config-
urations are most likely produced when OH is the oxidizing agent. When O and OH both exist, fully oxidized silicene may be
readily produced. Depending on the bonding of O and OH, the partially and fully oxidized silicene could be metallic,
semimetallic, semiconducting and insulating.
Recently, comprehensive studies of the oxidation processes on silicene in different phases have been reported by Du’s
group [40,182,336]. The correlation between buckled silicene surface and oxygen adatoms has been identified by using a
low-temperature STM and in-situ Raman spectroscopy. The electronic structures of oxidized silicene have been revealed
by ARPES as well as XPS. The formation of the Si–O bonds on surface of silicene was investigated under well-controlled
oxidation conditions with various concentrations of oxygen.
p p p p
The STM images of oxidized silicene in 4  4, 2 3  2 3, and 13  13 structures are shown in Fig. 65. It was found
that the initial oxygen absorption behavior (e.g., the initial absorption sites for oxygen adatoms) on silicene monolayers in
p p p p
different phases is dominated by their buckling structures. Si–O bonds dominate the 13  13 and 2 3  2 3 phases,
whereas Si–O–Si is the major oxidation form on 4  4 phase during the initial oxidation stage. In the former case, oxygen
adatoms occupy the top sites, while during oxidation of 4  4 phase the oxygen adatoms prefer to reside at the bridge sites.
The silicene oxides are distinct to silicon oxide. With increment of oxygen dose, silicene monolayers are eventually fully
oxidized due to the nature of mixed sp2/sp3 hybridization. The binding energy between epitaxial silicene layer and Ag
(1 1 1) surface is much smaller than the binding energy for Si–O. Therefore, the oxygen adatoms prefer to bond firstly with
Si atoms in silicene instead of Ag atoms in the substrate. Due to the characteristic sp3 hybridization of Si, energetically stable
Si–O–Si bonds would be expected with regardless of different phases, when silicene is exposed to high oxygen dose, for
example, 600 L. Interestingly, it was found that silicene monolayer crumples during oxidation and results in some ‘‘silicene
free” areas. These results suggest that, by taking this advantage, silicene oxide layer could be possibly detached from Ag
(1 1 1) substrate and become quasi freestanding.
The electronic structures of oxidized silicene have been investigated by STS, ARPES and DFT calculations. The ARPES
results are shown in Fig. 66. Before oxidation, the saddle-shape metallic hybrid surface band (HSB) can be observed in
epitaxial monolayer silicene, which is attributed to a hybridization of Si and Ag orbitals that resembles the p-band dispersion
in graphene. The saddle point of the surface state is about 0.15 eV below the Fermi level. ARPES investigation reveals that the
Ó American Physical Society 2016

Fig. 67. Possible adsorption sites (red), hollow, top, valley, and bridge, on a silicene lattice (blue). Reproduced with permission from
Ref. [134].
88 J. Zhao et al. / Progress in Materials Science 83 (2016) 24–151

Table 5
Calculated values for single adatom adsorption on silicene: relaxation sites hollow (H), bridge (B), valley (V) or top (T), adatom height (h), adsorption energy of
adatom (Eads), total magnetic moment of the system (l) in units of Bohr magneton (lB), energy band gap (Eg) and Bader charge transferred from adatom to
silicene (qad). Metallic and half-metallic systems are denoted as M and HM, respectively.

Site h (Å) Eads (eV) l (lB) Eg (eV) qad (e)


Li [134] H 1.69 2.40 0.0 M 0.8
Na [134] H 2.19 1.85 0.0 M 0.8
K [134] H 2.70 2.11 0.0 M 0.8
Be [134] V 0.78 2.87 0.0 0.39 1.3
Mg [134] V 1.98 1.22 0.0 0.48 1.0
Ca [134] B 1.49 2.68 0.0 0.17 1.3
Ti [134] B 0.77 4.89 2.0 HM 0.9
V [134] B 0.62 4.32 2.7 0.06 0.7
Cr [134] H 0.48 3.20 4.0 HM 0.3
Mn [134] H 1.04 3.48 3.0 0.24 0.4
Fe [134] H 0.33 4.79 2.0 0.18 0.0
Co [134] H 0.72 5.61 1.0 M 0.0
Ni [133] H 0.27 4.78 0.0 HM –
Mo [134] B 0.37 5.46 0.0 M 0.1
Pd [133] H 0.6 4.20 0.0 HM –
Ag [340] H 1.38 1.90 0.0 M 0.75
W [134] B 0.01 7.05 0.0 0.02 0.2
Pt [133] H 0.43 5.87 0.0 HM –
Au [133] H 1.01 2.32 0.57 M –
Hg [338] H 2.90 0.18 0.0 0.008 –
Al [341] V 1.71 2.87 0.0 M 1.0
Ga [133] V 1.70 2.40 0.0 – –
In [133] V 2.03 2.14 0.0 – –
Tl [338] H 2.06 2.17 0.92 M –
Sn [133] V 1.67 2.94 2.0 M –
Pb [338] H 1.61 2.64 0.0 M –
B [341] V 0.71 5.85 0.0 M 1.6
N [341] B 1.42 5.54 0.0 M 2.0
P [341] T 1.18 5.28 0.5 M 0.9

metallic HSB is decayed upon oxidation in silicene with an oxygen dose of 600 L. Meanwhile, the well-defined Shockley
surface state (SSS) and Ag-sp bands revive in the oxidized silicene/Ag(1 1 1) sample. Moreover, an asymmetric band with
the highest energy level of about 0.6 eV is observed in ARPES features, which suggests a formation of energy gap due to
oxidation.
STS measurements were performed on areas of Ag(1 1 1) surface and silicene oxide in the same sample. An energy gap of
1.2 eV is opened on silicene oxide, while Ag(1 1 1) shows a typical metallic characteristics. Therefore, the asymmetric state is
attributed to the valence band originating from silicene oxide. These results agree with the STM observations that
amorphous silicon oxide covers most area of the Ag(1 1 1) surface. It is worth to note that SSS in metal surface is extremely
surface sensitive that can reflect modifications of surface atomic and electronic properties. In the ARPES result, revived SSS in
oxidized sample demonstrates that oxygen would preferentially react with Si but not Ag when low oxygen dose is present.
As a result, Si–O bonds easily replace the Ag–Si bonds.

4.4. Metal adsorption

Due to highly reactive surface, the electronic properties of silicene can be easily tailored by adsorption of various foreign
atoms, including alkali metal, alkaline earth metal, transition metal (TM) and some main group elements [34,35,132–134,
331,332,337–346]. Fig. 67 displays all possible adsorption sites on a buckled silicene lattice, i.e., above the center of hexag-
onal rings (hollow site), on top of the upper atoms (top site), on top of the lower atoms (valley site), on top of the center of
Si–Si bond (bridge site). The favorite adsorption site, adatom height, adsorption energy, total magnetic moment, energy band
gap and Bader charge transferred from adatom to silicene for silicene sheets with various adatoms are summarized in
Table 5.
The adsorption configurations for different metal adsorbates on silicene are depicted in Fig. 68. For single alkali metal
atom on silicene, the favorite adsorption site is always the hollow site [133,134]. Upon relaxation, alkali metal atom sits
stably above the hollow site, staying about 1.69–2.70 Å away from the silicene surface (see details in Table 5) [134]. For
the same group of alkali metal elements, the distance between adatom and silicene surface monotonically increases with
increasing atomic size of the adatom. Adsorption of alkali atoms does not induce any significant distortion or stress on
the silicene lattice. The calculated adsorption energies of Li, Na, and K atoms on silicene are 2.40, 1.85, and 2.11 eV, respec-
tively [134]. However, the adsorption configuration may change with the concentration of alkali metal atoms. Osborn et al.
reports that the most stable adsorption site for Li on silicene would switch from hollow site to valley site when the adsorp-
tion ratio of Li atoms (defined as the number of Li atoms over the number of Si atoms) varies from 12.5% to 100% [337].
J. Zhao et al. / Progress in Materials Science 83 (2016) 24–151 89

Ó American Physical Society 2016

Fig. 68. Side (upper) and top (lower) views for adsorption geometries of alkali, alkaline earth, and transition metal atoms (yellow balls) on silicene (blue
balls). Reproduced with permission from Ref. [134].

Unlike alkali metals, hollow site is not the most favorable position for alkaline earth metals. Be and Mg favor the valley
site; while Ca prefers the bridge site [134]. Except for Mg, the adsorption energies of alkaline earth metals are slightly higher
than those of alkali metals (Table 5). When Be (Mg) monomer is adsorbed on the valley site of a silicene sheet, the original Si
atom would be pressed down (Fig. 68). Differently, adsorption of Ca atom tends to destroy the silicene lattice.
In contrast to alkali and alkaline earth metals, transition metal adatoms interact more strongly with silicene sheets except
for Ag, Au and Hg. In particular, the adsorption energy for Hg on silicene is only 0.18 eV, indicating a rather weak interaction
[338]. According to the DFT calculations by Sahin and Peeters [134], the most stable adsorption position is the bridge site for
Ti, V, Mo and W, which tend to destroy the silicene lattice. For the other transition metal adatoms (Cr, Mn, Fe, Co, Ni, Pd, Ag,
Pt, Au and Hg), the favorite site is the hollow site. As for the group-III and group-V elements, Al, Ga, In, and Sn favor the valley
site; Tl and Pb favor the hollow site. Non-metal atoms, i.e., B, N and P, have very large adsorption energy (>5 eV), implying
strong covalent bonding between adsorbates and silicene. However, Ni [133] and Gürel [339] reported that the favorite
adsorption position for Ti is the hollow site rather than the bridge site, while Ni [133] also proposed that Fe favors the valley
site, which is different from the finding by Sahin [134]. As listed in Table 5, most metal adatoms (except Fe and Co) donate
electrons to silicene, whereas nonmetal B, N and P atoms gain electrons form silicene. Generally speaking, this trend is a
consequence of electronegativity difference between silicon and the adsorbate elements.
Interestingly, the band gap of silicene can be opened by surface adsorption of alkali or transition metal atoms, by virtue of
a mixture mechanism of inversion symmetry breaking and intervalley interaction. It was found that alkali metal adsorption
90 J. Zhao et al. / Progress in Materials Science 83 (2016) 24–151

Ó American Physical Society 2016


Fig. 69. Electronic band structures for pristine, Li, Na, and K adsorbed silicene. Fermi level is set to zero. Occupied bands are filled with yellow color.
Reproduced with permission from Ref. [134].

can induce a small band gap at the Dirac point at low coverage (i.e., 31 meV for silicene + Li, 13 meV for silicene + Na, and
6 meV for silicene + K from the LDA calculations), nearly preserving the linear dispersion (Fig. 69) [134]. However, the entire
systems exhibit metallic character due to the charge transfer (about 0.8 electrons) from each metal atom to silicene, which
lifts up the Fermi level into the conduction band. It is worth to note that the interaction between the alkali atoms and silicene
is featured by ionic character, which wound not destruct the Dirac cone of silicene too much, in sharp contrast to the
covalent interaction between silicene and Ag substrates that destroys the Dirac cone and degrades of the carrier mobility
in silicene (which is discussed in Section 3.1.7).
Alkalis decorated silicene with different coverage were further investigated by Lu’s group [34]. As shown in Fig. 70, the
band gap increases with increasing coverage of Na adsorption. This can be understood by the following picture. When alkali
metal atoms are adsorbed on silicene, a vertical electric field is established because of the charge transfer from metal to
silicene. The electric field breaks the symmetry of the two sublattices in silicene and hence introduces a band gap. As the
coverage increases, the amplitude of electric field increases and the band gap is further enlarged owing to the enhanced
asymmetry between the two sublattices in silicene.
When the Li coverage is big enough and reaches one Li atom per Si, a fully lithiated silicene (namely, silicel) structurally
identical to silicane emerges, which is a direct semiconductor with band gap of 0.368 eV from the GGA calculation [337]. The
thermal and dynamical stabilities of silicel were confirmed by molecular dynamics simulation at 900 K over 3 ps and phonon
spectrum calculation. Recent DFT calculations show that small organic molecules (acetone) would enhance the adsorption of
Li atoms on silicene [342]. Moreover, a gap of 0.27 eV below the Fermi level by 0.06 eV is obtained in the case of Li doping on
acetone molecule adsorbed silicene, indicating potential application of this material in molecule based Li-ion batteries.
Ó Macmillan Publishers Limited 2016

Fig. 70. Na atom (purple) adsorbed silicene (yellow) layer. (a) Optimized geometry, (b) band structures. The Na atom coverage is 16.7% and 50.0%
respectively. The Fermi level is set to zero. The inset in the bottom of (b) is orbital of the conduction band bottom at the C point. (c) The k-space of the AMSi6
p p
supercell (one Na atom on ( 3  3) silicene). The Brillouin zones of the pure silicene and the AMSi6 monolayer are colored purple and green respectively.
Reproduced with permission from Ref. [34].
J. Zhao et al. / Progress in Materials Science 83 (2016) 24–151 91

p p
The intervalley scattering also plays a role to open the band gap when the supercell have a form of ( 3  3)n or (3  3)n
(n is an integer). In this case, the period in the BZ is shrinked comparing with the primitive BZ of silicene. For example, the
p p
reduced BZ (green color) of the Na-adsorbed silicene ( 3  3 silicene supercell) is shown in Fig. 70c, and that of primitive
silicene is colored purple. In the reduced BZ, the two inequivalent Dirac points of silicene, Ks and Ks0 , are mapped on the C
point through the reciprocal unit vector G1 or G2 (an extra p/6 rotation with respect to the original BZ of silicene also takes
places here). The interaction between the two valleys Ks and Ks0 further enlarges the band gap in silicene. Benefited from the
gap opening by alkali metal adsorption, silicene based FET can be constructed, which will be discussed later in Section 5.1.3.
As for the effect of other main group elements, Be, Mg and Ca adsorption on silicene would open band gaps of 0.39 eV,
0.48 eV and 0.17 eV, respectively, transforming silicene into a semiconductor [134]. No magnetism is induced by adsorption
of either alkali or alkaline earth metal. The adsorption of non-metal elements (B, N and P) would transform silicene from
half-metal to metal [341]. No magnetism would be induced by B and N adsorption, while P adsorption would introduce a
magnetic moment of 0.5 lB per P atom. Valley site is the most stable adsorption site for all Group III metal adatoms
(Al [341], Ga [133] and In [133]), except that Tl favors the hollow site [338]. Moreover, Tl can induce a magnetic moment
of 0.92 lB, while no magnetism would be brought in by adsorption of Al, Ga and In atoms. As for group IV metal adatoms,
Sn prefers the valley site [133], while Pb favors the hollow site [338]. Sn and Pb decorated silicene sheets are ferromagnetic
metal and nonmagnetic metal, respectively [133,338].
Adsorption of transition metal atom may also open a sizeable band gap at the Dirac point of silicene without degrading its
electronic properties [33]. The size of the gap ranges between 0.03 and 0.66 eV, depending on type and coverage of TM
dopants. Different from alkali metal atom adsorption, which only induces n-type doping in silicene due to their smaller work
functions, transition metal metals induce different doping type, depending on their work functions (n-type by Cu, Ag and Au
adsorption, p-type by Ir adsorption, and neutral type by Pt adsorption). Even though the charge transfer between Pt atoms
and silicene is negligible, the wave function interaction between Pt and the bottom Si layer breaks the inversion symmetry of
silicene and opens the band gap in Pt-adsorbed silicene. When the two valleys in transition metal atom adsorbed silicene are
folded together, the intervalley interaction greatly enhances the band gap.
In addition to the effect of gap opening, adsorption of most transition metal atoms might also introduce abundant
magnetic properties and topological states in silicene. Hence, the transition metal decorated silicene materials are attractive
due to the potential applications in spintronics and magnetic storage. The adsorption of a Ti, V, Cr, Mn, Fe, Co and Au atom on
a silicene sheet would result in ferromagnetic ground states with net magnetic moments of 2.0 lB, 2.7 lB, 4.0 lB, 3.0 lB,
2.0 lB, 1.0 lB, 0.57 lB per supercell, respectively [133,134].

Ó American Institute of Physics 2016

Fig. 71. (Left panel) l-LEED patterns recorded at 30 eV, corresponding to the K atom coverage of: (a) x = 0; (b) x = 0.08 ± 0.01; and (c) x = 0.16 ± 0.03. (Right
panel) ARPES spectra of epitaxial silicene on ZrB2(0 0 0 1) along the C—M—C direction, in the vicinity of the M point of the 1st BZ: (a) x = 0; (b) x = 0.09 ± 0.01;
and (c) x = 0.18 ± 0.03. Reproduced with permission from Ref. [350].
92 J. Zhao et al. / Progress in Materials Science 83 (2016) 24–151

Recently, the magnetic properties of silicene with Ti adsorbed on the hollow site was explored by Gürel and co-workers
[339]. They found that charging and perpendicular electric field can affect the magnetic moment of Ti adsorbed silicene and
an electric field of 1 V/Å can even switch the ground state from ferromagnetic to antiferromagnetic. Magnetic moments of
3.50 lB and 4.56 lB have been obtained for Mn decoration at the top and hollow sites of BN supported silicene, respectively
[343]. Interestingly, the orientation between the Mn moment and the induced Si moments is ferromagnetic in the former
and antiferromagnetic in the latter case. Furthermore, Ni’s group revealed highly tunable magnetism (up to about 4 lB
per TM atom) in Cr and Fe doped silicene monolayers under isotropic and uniaxial tensile strain [346].
When the center of each hexagonal ring in silicene is filled with one transition metal atom, a series of novel 2D materials,
namely, transition metal silicide (TMSi2) monolayers, occur. Recently, twenty TMSi2 monolayers, including all 3d and 4d
transition metal elements, have been studied by first-principles calculations [347]. Among them, TiSi2, VSi2, CrSi2, NbSi2
and MoSi2 are ferromagnetic, MnSi2 and FeSi2 monolayers adopt antiferromagnetic; while all the rest of TMSi2 monolayers
are nonmagnetic. The on-site moments of TM atoms in TMSi2 (TM = Ti, V, Cr, Nb, Mo, Mn and Fe) monolayers are 0.563 lB,
2.148 lB, 3.008 lB, 0.618 lB, 0.207 lB, 2.512 lB and 1.058 lB, respectively. Especially, the formation energies of TiSi2, FeSi2
and MnSi2 monolayers are lower than or equal to that of silicene, indicating their possible synthesis in experiment.
Interestingly, in an independent theoretical study, CrSi2 monolayer was predicted to exhibit a strong piezomagnetic
coupling [348].
Recently, the structural stability and magnetic properties of TM atoms from Sc to Zn embedded into silicene with single
vacancy (SV) and double vacancies (DV) were investigated by first-principles calculations [349]. The TM atoms bond strongly
with the surrounding Si atoms for both SV and DV with binding energy larger than 4 eV, except for Zn in which case the bind-
ing energy is about 2 eV. Large magnetic moment, which mainly comes from the 3d orbitals of transition metal along with
partly contributions from the neighboring Si atoms, can be induced in silicene with SV or DV by embedding TM atoms from V
to Co. While, Sc, Ti, Ni, Cu and Zn decorated silicene sheets with SV or DV show nonmagnetic ground state. Moreover, the
magnetism of the TM atoms embedded silicene can be greatly enhanced by N or C doping at SV site in silicene.
Owing to the magnetic doping and strong spin–orbit coupling in silicene, transition metal atom adsorbed silicene may
provide abundant topological states. Through first-principles calculations, Kaloni and co-workers demonstrated that
Co-decorated silicene can host a quantum anomalous Hall state [344]. To consider the concentration effect of transition
metal atoms, they put one Co atom in 1  1, 2  2, 3  3 and 4  4 supercells of silicene, respectively. They found that
QAH state can only exist for one Co atom in a 4  4 supercell, indicating that the interaction between transition metal atoms
counteracts the creation of this state and that the impurity density of TM atoms has to be sufficiently low. Recently, Zhang

Ó IOP Publishing 2016

Fig. 72. (a) Geometry and (b–f) band structures of alkali metal intercalated bilayer silicene. The yellow balls and purple balls represent Si and alkali metal
atoms, respectively. Reproduced with permission from Ref. [35].
J. Zhao et al. / Progress in Materials Science 83 (2016) 24–151 93

et al. proposed that V doped silicene is in QAHE(+2) phase and may exhibit QVHE(0) phase or QAHE(2) phase under certain
external electric field [132]. Moreover, systematical investigation of the topological properties of silicene adsorbed with 4d
TM atoms was conducted by Yang’s group [345]. Two kinds of QAH states with Chern numbers of 2 and 1 were found in
Nb/Ru and Y doped systems, respectively. The other topological states of valley Hall effect (VHE) and valley-polarized metal
(VPM) were obtained in Y/Mo-silicene and Zr/Tc-silicene, respectively. These exotic topological states may be destroyed if
the TM coverage is higher than 5%.
Parallel to the above mentioned theoretical efforts, aluminum and potassium atoms have been deposited on silicene
sheets in experiments. After aluminum ultrathin film of few nms thickness forms on Ag(1 1 1)-supported silicene sheet, a
shift of the Si 2p XPS line towards low binding energy was observed [331]. This can be related to the hybridization between
Al and Si orbitals, as theoretically predicted by Lin and Ni [133]. Meanwhile, Friedlein et al. [332] deposited monolayer of Al
atoms on epitaxial silicene on ZrB2(0 0 0 1) surface up to 0.38 ± 0.06 Al atoms per Si atom and found that the effects of Al
deposition on the Si 2p spectra are rather small, indicating a weak interaction between the Al atoms and silicene.
Using ARPES and microspot LEED (l-LEED) techniques, Friedlein et al. studied the evolution of the electronic structure
and the structural stability of various coverage (x) of K atoms adsorbed on silicene monolayer grown on thin ZrB2 films
[350]. First, adsorption of K atoms leads to reduction of work function (D/), e.g., D/ = 1.2 eV for x = 0.2, and D/
p p
= 1.5 eV for x = 0.55. As shown in Fig. 71, the 3  3 reconstruction of silicene does not change upon exposure to potas-
sium atoms up to a level of x = 0.16, indicating that the silicene layer remains structurally intact maintaining the epitaxial
relation with the diboride surface and the degree of buckling of Si atoms does not change much. Upon deposition of K atoms,
a band denoted ‘‘X5” in ARPES spectrum appears at EF right at the M point. This new electron pocket is formed by electron
donation from the alkali metal atoms to a formerly unoccupied (likely p⁄) band. K adsorption also results in a shift of
p-electronic states towards higher binding energy and an increasing hybridization between Zr 4d- and Si p-derived
electronic states.

Fig. 73. (a) High-resolution ARPES spectra near the Dirac point (ED) at around K(H) point in tr3-CaSi2. (b) ARPES-intensity plot around K(H) point as a
function of wave vector and binding energy. Peak position in the ARPES spectra obtained by fitting with Lorentzians is shown by blue and red circles. (c)
Second-derivative ARPES intensity plot around ED. (d–f) ARPES intensity mappings in a 2D vector plane at three representative energy slices, ED + 0.3 eV, ED,
and ED  0.3 eV, respectively. Reproduced with permission from Ref. [351]. Copyright 1999–2016, John Wiley & Sons, Inc.
94 J. Zhao et al. / Progress in Materials Science 83 (2016) 24–151

Besides the adsorption of metal atoms on monolayer silicene, intercalation of alkali metal atoms in bilayer silicene has
been theoretically proposed by Zhao’s group using first-principles calculations [35]. The geometry of alkali metal
intercalated bilayer silicene (named as LiSi4, NaSi4, KSi4 and RbSi4) is shown in Fig. 72a. Strong coupling between the two
layers in bilayer silicene is removed by intercalation, as evidenced by the large inter-layer distance (d = 3.884 Å for Li,
d = 5.892 Å for Rb) and the buckling height (h = 0.659 Å for Li, d = 0.504 Å for Rb) close to the freestanding silicene
(h = 0.467 Å). The formation energies for most metal intercalated bilayer silicene sheets (LiSi4, NaSi4 and KSi4) are only about
0.3 eV per atom, which are much lower than that for freestanding monolayer silicene (0.779 eV per atom). This suggests that
alkali metal intercalation can significantly improve the energetic stability of silicene. The thermal stability of KSi4 was
assessed by AIMD simulation with temperature of 500 K and time step of 1.0 fs. No topological defect in silicene lattice
was ever generated during the entire simulation time of 10 ps, confirming the high thermal stability of K-intercalated bilayer
silicene.
As shown in Fig. 72, the band structures for all intercalated systems more or less resemble that of monolayer pristine
silicene. For LiSi4 (Fig. 72b), a pair of Dirac-like bands appears below Fermi level in the C–K region and the M point in the
first BZ. For NaSi4 (Fig. 72c), the two Dirac-like bands move closer to the K point. Eventually, these two Dirac-like bands
coincide at the K point for KSi4 (Fig. 72d), forming a double-degenerated Dirac cone with a band gap of 0.27 eV. In the case
of RbSi4 (Fig. 72e), the two Dirac-like bands move over the K point and split again. However, KSi4 is metallic due to the charge
transfer from K atoms to silicene. Alternatively, the band structure of K+Si4 (Fig. 72f) exhibits distinct semiconducting
character with a moderate gap of 0.43 eV and relatively small effective masses of hole (0.174 me) and electron
(0.215 me). In a FET, the charge doping can be realized by applying a suitable gate voltage. Thus, K+Si4 would be feasible
under a proper bias voltage perpendicular to the silicene sheet in a FET device.
Soon after the theoretical proposal of metal intercalated silicene, Takahashi’s group reported an ARPES study of calcium
disilicide (CaSi2), which can be regarded Ca-intercalated multilayer silicene [351]. In the ARPES-derived band structure, they
observed the massless Dirac-cone p-electron dispersions at the K(H) point in the Brillouin zone (Fig. 73), together with the
r-band dispersions at the C point. Due to the substantial charge transfer from Ca atoms to silicene layer, the Dirac point is
located far away from the Fermi level by about 2 eV. The experimental realization of CaSi2 paves the way of restoring the
Dirac dispersion in layered materials by means of metal intercalation. Later, the electronic properties of the silicene-
intercalated compound CaSi2 have been explored with ab initio calculations by considering several possible stacking
sequences [352]. It was found that the Dirac cone shifts away from the high symmetric point of the hexagonal Brillouin zone

Fig. 74. Atomic structures for (a) pristine zigzag nanoribbon of w = 6, (b) pristine armchair nanoribbon of w = 8, (c) (2  1) reconstructed zigzag nanoribbon,
(d) (2  1) reconstructed armchair nanoribbon of w = 9. The lines denote the unit cells of 2D lattice.
J. Zhao et al. / Progress in Materials Science 83 (2016) 24–151 95

with the opening of a small gap. This was attributed to the sublattice symmetry breaking and consequent enhanced asym-
metric interlayer hopping, enforced by the enhanced buckling of the silicene layers.

4.5. Confinement effect

The realistic device applications of silicene may involve small pieces of 2D sheets rather than infinite monolayer mate-
rials. In this regard, quantum confinement effect due to finite size would be expected and there are additional degrees of
freedom from the edge atoms. Therefore, the electronic properties of finite-width 1D silicene sheets (namely, nanoribbon)
or finite-size 0D silicene clusters (or nanoflakes) have been intensively studied. The experimental synthesis and spectro-
scopic characterization of silicene nanoribbons have already been described in Section 3.3. Here we focus on the theoretical
calculations of silicene nanoribbons (NRs), nanoflakes, and their hydrogenated or fluorinated derivatives [8,41,353–371]. The
spintronics with silicene nanoribbons will be discussed in Section 5.3.1.
Usually, two kinds of silicene nanoribbons (namely, armchair and zigzag) have been considered. As shown in Fig. 74, the
width of nanoribbons (w) is defined as the number of Si atoms forming a continuous chain between the two edges. For all
bare nanoribbons without hydrogen passivation on the edge Si atoms, edge reconstruction occurs upon relaxation; whereas
reconstruction disappears when the edge dangling bonds are fully terminated by hydrogen atoms. For the bare armchair
NRs, a (2  1) asymmetric dimer-like reconstruction at the edges was found [8,353], as displayed in Fig. 74d. In the case
of zigzag NRs, a new (2  1) reconstructed edge configuration (Fig. 74c) composed of a triangle–pentagon pair topological
defect was found to be very stable with a low edge energy of 0.32 eV/Å [354,355].
With hydrogen passivation, all armchair silicene nanoribbons (ASiNRs) are nonmagnetic semiconductors with band gaps
relative to the width w. As shown in Fig. 75, the variation of band gap with w exhibits an oscillation: if w = 3p + 2 (p being an
integer), band gap is very small [8] or zero [41,356]; band gap is large for w = 3p and w = 3p + 1 (the latter being largest for a
given p). Fig. 75 displays the oscillatory decrease of band gap with the increase of nanoribbon width [41]. First-principles
calculations by Song et al. [357] further demonstrated that external transverse electronic fields can significantly modulate
the band gap of an armchair nanoribbon, which can be explained by the Stark quadratic effect and have potential applica-
tions in nanoelectronic devices.
For the zigzag silicene nanoribbons (ZSiNRs) with hydrogen termination on the edge, magnetic states are energetically
favorable over the nonmagnetic states, and the ferromagnetic and antiferromagnetic states are nearly degenerate in energy
[8,41]. As the ground state configuration that is slightly more stable, the antiferromagnetic silicene zigzag nanoribbons are
semiconductors with band gaps decreasing gradually as the ribbon width increases (Fig. 75). The edge states have opposite
spins and approximately zero total magnetic moment within the entire unit cell [41]. Under transverse electric field,
the band gaps of up-spin and down-spin change in different manners; consequently, the zigzag silicene NRs become

Ó American Institute of Physics 2016

Fig. 75. (a) Variation of band gap with ribbon width w for hydrogen saturated armchair ribbon. (b) The spin density distribution of hydrogen saturated 6-
ZSiNRs at antiferromagnetic state. (c) Variation of band gap with ribbon width w for hydrogen saturated zigzag ribbon. Reproduced with permission from
Ref. [41].
96 J. Zhao et al. / Progress in Materials Science 83 (2016) 24–151

half-metals. Meanwhile, perpendicular electric field can also modulate the energies of spin-polarized edge states and the
corresponding band gaps, due to the localized edge states and the buckled structure of silicene zigzag NRs [358]. This
modulation effect is width dependent and becomes more significant for the narrower NRs.
Using a model TB Hamiltonian with inclusion of the spin–orbit interaction, Ezawa and Nagaosa [359] investigated the
electronic band structures of armchair and zigzag silicene nanoribbons up to width of w = 100. Due to finite size effect,
the gap 2d of the edge channel for armchair NRs increases with decreasing w. When the width is 10 lm, the gap 2d is in
the order of 10 meV, which is experimentally observable. Different from the graphene nanoribbons, the edge channel in
silence NRs is topologically protected from disorder as long as the time-reversal symmetry is preserved and the bulk gap
remains finite. The same group has further studied the edge states of silicene nanodisks and observed helical currents
flowing around the edge, because silicene is a topological insulator due to the spin–orbit interaction [360].
In addition to the regular zigzag and armchair edges, Zhao and Ni investigated the relative stability and electronic prop-
erties of silicene nanoribbons with sawtooth edges (SSiNRs) and edge Si atoms being passivated by H atoms [361]. Their first-
principles calculations showed that the SSiNR is more stable than the zigzag silicene nanoribbon and has a ferromagnetic
ground state with an intrinsic energy gap between majority and minority spin-polarized bands (namely, a spin-
semiconductor), which originates from the edge states. The electronic structure of SSiNRs can be modulated by external
electric field and strain. Under a relatively low transverse electric field (0.05 eV/Å) or suitable compressive strain
(4.6%), the SSiNRs can be converted into a spin gapless semiconductor.
In the theoretical calculations, the edges of Si NRs are usually fully passivated by hydrogen with sp2 hybridization
(i.e., one H atom for one edge Si atom). Alternatively, Ding and Wang [362] considered zigzag silicene NRs with asymmetric
sp2–sp3 edges (Fig. 76a), i.e., each Si atom on one side terminated by one H atom (sp2), and each Si atom on another side
terminated by two H atoms (sp3). As depicted in Fig. 76b, the di-hydrogen termination almost eliminates the atomic moment
of edge Si1 atom (only 0.02 lB). The Si2 and Si12 atoms with largest moments belong to the same sublattice of nanoribbons
and exhibit ferromagnetic coupling. Band structures in Fig. 76c manifest that the asymmetric 6-ZSiNR is a ferromagnetic
semiconductor with an indirect band gap of 0.22 eV. Around the Fermi level, the states are completely spin-polarized with
opposite spin orientations, declaring the bipolar magnetic feature. These asymmetric ZSiNRs can be further modified to half
metals by substitutional doping of B and P atoms, leading to opposite conductive spin channels.
Moreover, Zhang’s group [363] explored the effect of dangling bonds on the electronic and magnetic properties of both
zigzag and armchair nanoribbons. They found that the zigzag nanoribbons with either one or two bare sides are antiferro-
magnetic, compared to the ferromagnetic state of the perfect ZSiNRs. Although the perfect ASiNRs with hydrogen termina-
tion are non-magnetic, the dangling bonds at one edge and both edges result in ferromagnetic and antiferromagnetic states,
respectively. For the silicene NRs with ferromagnetic state, successive removal of hydrogen atoms on the edge would
enhance the magnetic moment, e.g., from 0.43 lB per unit cell of perfect 6-ZSiNR to 0.58 lB of one bare side, and to
0.75 lB of two bare sides. However, Zhang’s calculations did not account for the possible edge reconstruction [8,353–
355]. With (2  1) reconstructed edges, the bare armchair NRs are still non-magnetic semiconductors and exhibit width-
dependent oscillation of band gap similar to the H-passivated ones. On the other hand, the antiferromagnetic and ferromag-
netic states of bare zigzag NRs with reconstructed edges are nearly degenerate in energy (the former being slightly more
stable by DE = 4 meV) [354].
Beyond pristine silicene NRs, the completely hydrogenated or fluorinated silicene NRs, namely, H-silicane or F-silicane
nanoribbons with chair-like configurations have been theoretically explored by Fang and co-workers [364]. Although the
Ó American Institute of Physics 2016

Fig. 76. (a) The atomic structure, (b) distributions of atomic magnetic moments, (c) spin-polarized band structures for the asymmetric 6-ZsiNRs.
Reproduced with permission from Ref. [362].
J. Zhao et al. / Progress in Materials Science 83 (2016) 24–151 97

Fig. 77. Geometries of various point defects, nanomesh and extended line defect in silicene: (a) SW; (b) SV-1 by (55|66) rings; (c) SV-2 with three dangling
atoms; (d) DV-1(5|8|5); (e) reconstructed DV-2(555|777); (f) TV; (g) Si adatom (highlighted by blue, sideview is shown on top); (h) nanomesh with
hydrogen atoms (white ball) terminating the dangling Si atoms; (i) 5|8|5 line defect.

armchair H-silicane NRs exhibits an indirect band gap like the infinite H-silicane sheet, the zigzag H-silicane NRs and all F-
silicane NRs are direct-band-gap semiconductors. For both H-silicane and F-silicane, the armchair NRs have a slightly larger
band gap that the zigzag ones, and their band gap decreases as the ribbon width increases. The band gaps of
F-silicane NRs (around 1 eV) are about half of those of H-silicane NRs (c.a. 2.526–2.410 eV). The electronic structures of these
silicane nanoribbons can be further modulated by a transverse electric field, i.e., making the electron states and hole states
spatially separated and rendering silicane NRs self-doped (p-type on one side and n-type on the opposite side). Recently,
Jiang et al. considered the hybrid silicene/fluorinated silicene nanoribbons (SFNRs) with both zigzag and armchair interfaces
[372]. The armchair SFNRs are nonmagnetic semiconductors with oscillating band gap as function of the width of silicene
part, while the zigzag SFNRs are antiferromagnetic semiconductors.
Using first-principles calculations, Zhang’s group also explored the electronic and magnetic properties of silicene nano-
flakes of hexagonal and triangular shapes with hydrogen-passivated zigzag edges [365]. They found that the hexagonal sil-
icene nanoflakes are non-magnetic semiconductor with size-dependent band gap due to quantum confinement effect of
electron motion from 3p orbitals of Si. In contrast, the triangular nanoflakes exhibit magnetic behavior with integer magnetic
moment of m–2 lB, where m is number of zigzag lines. When two triangular nanoflakes are linked by a Si chain, ferromag-
netic coupling between two nanoflakes are found for the odd-numbered Si chain, whereas directly linking two nanoflakes
leads to antiferromagnetic coupling. In future experiment, it would be interesting to create finite-size silicene nanoflakes
intentionally and to characterize the electronic properties using spectroscopic means.
Using a simplified tight-binding Hamiltonian for the low-energy states, the electronic properties of triangular and hexag-
onal nanoscale quantum dots (QD) of silicene in transversal electric field have been discussed [373]. The edge-localized zero
energy states (ZES) are found in the middle of the size-quantized gap and can be easily tunable by electric field, suggesting
potential field-effect scalable QD devices.
98 J. Zhao et al. / Progress in Materials Science 83 (2016) 24–151

Ó Royal Society of Chemistry 2016

Fig. 78. Energy barriers for (a) transition from pristine silicene to SW defect, (b) diffusion of SV-1, (c) transition from DV-1(5|8|5) to DV-2(55|77), (d)
diffusion of DV-1. Structural images along the transition path are shown on the right panels. Reproduced with permission from Ref. [377].

4.6. Defect effect

Like the case of graphene [374], defects are nearly inevitable during the growth of silicene [23,179,180,200,206–208,2
39,375,376]. Hence, how defects affect the physical properties of silicene becomes an important issue. The geometries of
some representative defects in monolayer silicene sheet, including Stone–Wales (SW) rotation, single-, double-, and
triple-vacancies (abbreviated as SV, DV, TV, respectively, hereafter), Si adatom, nanomesh (or antidot), and line defect are
presented in Fig. 77.
J. Zhao et al. / Progress in Materials Science 83 (2016) 24–151 99

Ó Royal Society of Chemistry 2016


Fig. 79. Electronic band structures of pristine and defective silicene: (a) pristine silicene (the Dirac point is highlighted by a red circle), (b) SW defect,
(c) SV-1(55|66), (d) DV-1(5|8|5), (e) DV-2(555|777) and (f) Si adatom. The Fermi level is set to zero and denoted by red dashed line. The red solid lines in (f)
stand for spin up, while blue solid lines for spin down. Reproduced with permission from Ref. [377].

4.6.1. Stone–Wales defects


As a common topological defect in the honeycomb lattices, Stone–Wales defect (Fig. 77a) is obtained by a 90 deg rotation
of a Si–Si bond, forming two pentagonal rings and two heptagonal rings: (5|7|7|5). For the infinite silicene sheet, the forma-
tion energy of a SW rotation from DFT calculations is 2.09 eV by Gao et al. [377], 1.64 eV by Sahin et al. [378], 1.82 eV by
Manjanath and Singh [379], 1.84 eV by Haldar et al. [380], respectively. Meanwhile, in a finite silicene flake with 80 Si atoms
and hydrogen termination on the edges, the formation energy of a SW defect decreases from center to edge, i.e., 1.36 eV
(center), 1.23 eV (towards center) and 1.16 eV (edge), which are all lower than the bulk value of 1.82 eV obtained using
the same computational method. Similar trend was observed in zigzag silicene nanoribbons by Dong and co-workers
[381]. For instance, for a 5-ZSiNR of 1.932 nm width, the formation energy for SW rotation is as low as 0.68 eV. Therefore,
the possibility of having such SW defects in finite silicene sheets is even higher than the infinite systems.
The energy barriers for the formation of a SW defect inside silicene sheet have been obtained by first-principles compu-
tations of the energy profile along the rotation path. For a freestanding sheet, the calculated energy barrier for structural
transition from pristine silicene to SW defect are 2.64 eV [377], 2.4 eV [378], 2.91 eV [379], respectively, see Fig. 78a. This
barrier is even lower in the finite-width nanoribbons, e.g., only 1.63 eV in the 5-ZSiNR [381]. Sahin et al. [378] further inves-
tigated SW rotation in a silicene sheet supported on Ag(1 1 1) substrate and found that the energy barrier raises by about
0.4 eV due to the interaction with substrate. Compared with the very high energy barrier for SW rotation in graphene
(10 eV) [374], SW defects can be more easily created in silicene. On the other hand, the reverse barrier from SW back to
perfect silicene is only 0.5 eV (Fig. 78a), suggesting that annealing at appropriate temperature may eliminate the existing
SW defects in a silicene sheet [377].
First-principles band structure calculations revealed that the existence of a SW defect in silicene would induce a tiny band
gap by breaking the hexagonal symmetry, e.g., 33 meV for (5  5) supercell [377], 20 meV for (6  6) supercell [378], 10 meV
for (7  7) supercell [379], respectively; though the specific amplitude of the opened gap vary with the details of computa-
tional method and size of supercell. Most importantly, the defective silicene sheet with SW defect still retains the linear
dispersion relationship near the Fermi level (Fig. 79a). In other words, a silicene lattice with a small amount of SW defects
would open a small gap without losing the high carrier velocity. This is beneficial for future applications in nanoelectronic
100 J. Zhao et al. / Progress in Materials Science 83 (2016) 24–151

devices. In a recent study, it was also shown that incorporation of a SW defect in armchair silicene nanoribbons only slightly
affect the I–V curve for a bias up to 1.0 V [382]. Similarly, Haldar et al. [380] also found that SW defect does not result in
considerable changes in the calculated transmission coefficients compared to the pristine silicene sheet apart from a small
decrease in the slope of the curve.

4.6.2. Vacancies
Vacancies with one to a few missing lattice atoms are another type of common structural defects in 2D honeycomb
crystals. As shown in Fig. 77, two kinds of single vacancies and two kinds of double vacancies within infinite silicene sheet
have been considered by Gao et al. [377]. After full DFT relaxation, SV-2 with three dangling atoms is less stable than SV-1
with (55|66) rings by 0.76 eV. Later, Dong’s group [383] considered other two possible SV configurations and still found that
SV-2 is the most stable one. In the case of DV, a reconstruction from DV1 with (5|8|5) rings into DV2 with (555|777) rings
lowers the formation energy by 0.86 eV and needs to overcome an energy barrier of 1.2 eV [377], as displayed in Fig. 78c.
Note that the formation energy for a DV (3.70 eV for DV1 or 2.84 eV for DV2) is much smaller than the summation of
two isolated SVs (e.g., 6.02 eV for two SV1), implying that two SVs may coalesce into a DV via diffusion. As depicted in
Fig. 78b, DFT calculations demonstrated that SV-1 defect can easily migrate inside silicene lattice by overcoming a small
energy barrier (0.12 eV [377] or 0.15 eV [383]) and the corresponding diffusion coefficient is in the order of 105 cm2/s. Thus,
combination of two SVs into a DV via diffusion is probable. In contrast, a large energy barrier of 2.06 eV would prevent the
migration of DV-2 (Fig. 78d). Using DFT calculations combined with AIMD simulation, Li et al. [384] recently considered
larger vacancy clusters with up to six missing atoms and demonstrated that the smaller vacancies in silicene tend to coalesce
and reconstruct into larger vacancy cluster to reduce the energy.
Using first-principles calculations, Ciraci’s group studied the self-healing mechanisms of various vacancy defects (i.e., sin-
gle, double, and triple vacancies) in silicene by considering atomistic processes of host adatom adsorption, diffusion, vacancy
formation [385]. For all three kinds of vacancy defects explored, bond reconstructions at defect site are exothermic
processes, which in turn become the driving force for the healing process. Under the external supply of host atoms, the
Stone–Wales defect can form in the course of healing of silicene vacancy, which can be further recovered by a bond rotation
as discussed in Section 4.6.1.
Defects can effectively modify the electronic properties of silicene. In the case of infinite silicene sheet modeled by a
(5  5) supercell (Fig. 79), SV-1(55|66) defect would transform semimetallic silicene to metallic, DV-1(5|8|5) defect would
introduce a moderate band gap of 161 meV in silicene, and DV-2(555|777) would destroy the Dirac cone while keep the
semimetal characteristic of silicene, respectively [377]. For an armchair silicene nanoribbon with 13 Si–Si dimer lines across
the ribbon width, namely, 13-ASiNR, it was found that either a SV or a parallel oriented DV transforms the direct semicon-
ductor into an indirect one, while a slanting oriented DV induces metallic characteristic [386]. Very recently, modification of
transmission coefficients in silicene sheet by SV2, DV-1(5|8|5), DV-2(555|777) defects was also theoretical explored by
Haldar et al. [380] using DFT and non-equilibrium Green’s function method.
In addition to the perturbation on electronic band structures, existence of vacancies in a monolayer silicene has certain
impacts on its thermal conductivity and thermal stability. Using equilibrium MD simulation combined with the Green–Kubo
formula, Li and Zhang [387] simulated the thermal conductivity of silicene sheets with various situations of vacancies. They
found that introducing vacancies and vacancy clusters significantly reduces the in-plane thermal conductivity of silicene
nanosheet owing to the strong effect of phonon-vacancy scattering. Such effect is more pronounced as the concentration
and size of vacancy defects increase, which can be attributed to the depression and broadening of some phonon vibration
modes. Moreover, the edge shape of vacancy clusters has a significant influence not only on the amplitude of thermal
conductivity, but also on its anisotropy. With a 960-atom supercell of silicene, MD simulation using ReaxFF predicted a
significant reduction of melting temperature from Tc = 1500 K for pristine silicene to Tc ffi 1000 K for silicene sheet with a
monovacancy [57]. It was found that Tc reduces almost linearly with further increasing the size of vacancy up to five missing
atoms.
Ciraci’s group [388] has predicted the increased chemical activity of dissociative adsorption of small molecules at the SV2
defect of silicene. Dissociations of H2, O2, CO molecules at monovacancy are all exothermic processes, with energy barriers of
1.30, 0.05, and 4.36 eV, respectively. Meanwhile, the H2O molecule is inert to defective silicene. Therefore, silicene can be
easily functionalized by creating meshes of a single vacancy.
In contrast to the high reactivity of SV, the porous silicene with DV-1 and DV-2 defects without additional passivation of
the dangling silicon atoms is chemically inert to noble gases (He, Ne and Ar), H2, H2O, CO, CO2, N2 and CH4 and thus shows
great prospect in the gas separation and filtering applications [389,390]. For example, the diffusion energy barrier for H2
penetrating through a DV-1 hole of silicene is only about 0.34 eV, suggesting that a H2 molecule can permeate the porous
silicene at moderate temperature and pressure [389]. On the other hand, the calculated diffusion barriers are as high as
1.03, 0.99, 1.01 and 1.66 eV for N2, CO, CO2 and CH4 molecules, respectively. Therefore, porous silicene sheet with
subnanometer pores would serve as a good hydrogen purification membrane [389]. Similarly, helium atom can pass through
the DV-1 and DV-2 defects of silicene easily with low energy barriers of 0.57 and 0.33 eV, respectively, exhibiting high
selectivity for He/Ne and He/Ar mixture gases [390].
Compared to the extensive theoretical efforts on the various structural defects in freestanding silicene, experimental
identification of all kind of defects in the as-prepared silicene samples on substrates remains a challenge. For this purpose,
one must consider the effect of metal substrates. Dong’s group investigated the stability and electronic properties of SV in
J. Zhao et al. / Progress in Materials Science 83 (2016) 24–151 101

4  4 silicene structure on Ag(1 1 1) substrate using first-principles calculations [391]. Very recently, Liu and co-workers
performed a systematical study on the morphology and stability of various point defects in epitaxial silicene on Ag(1 1 1)
surfaces using first-principles calculations combined with experimental STM observations [376]. The atomic structures
for some point defects observed in experiments were identified by the excellent accordance between their simulated and
p p
experimental STM images. The equilibrium concentrations of single and double vacancies in 13  13 silicene at 500 K
13 2 13 2
were estimated to be as high as 4.4  10 cm and 5.0  10 cm , respectively, which explains the defective appearance
p p
of 13  13 silicene in experiments [22,23,375,376,392]. Moreover, Si-SV can diffuse very fast in both 4  4 and
p p
13  13 silicene sheets at 500 K. Thus, two SVs would easily coalesce into one DV via diffusion to lower energy.

4.6.3. Nanomeshes
As a strategy to open the band gap in graphene around the Fermi level, graphene nanomeshes (or antidot lattices) have
been proposed [393,394]. By analogy, there have been some theoretical investigations on the electronic, optical, and
transport properties of silicene nanomeshes (SNMs) with periodic array of hexagonal or triangular holes within the pristine
silicene sheet, where the dangling bonds of silicon atoms at hole edge are usually passivated by hydrogen atoms [395–399].
It was found that the electronic and optical properties of silicene nanomeshes can be tuned by the width (W) of the silicon
chains between neighboring holes [395,397,399]. Moreover, the dielectric functions and optical absorption spectra of the
SNMs show a broad frequency photoresponse ranging from 0 eV to the ultra-violet region 10 eV depending on W, which
would be useful for harvest of solar energy.
As a presentative, the SNM model with simple triangular array of hexagonal holes is shown in Fig. 80a. The type of SNM is
notated by [R, W], where R is the radius of the hole calculated by Nremoved = 6R2 (Nremoved is the number of the removed Si
atoms from one lattice cell) and W is the width of the wall between the nearest-neighboring holes. When W is odd, the SNMs
behave semimetallically like pristine silicene (Fig. 80b). While W is even, the band gap at the U point is opened (Fig. 80c), e.g.,
0.64 eV for W = 2, and 0.32 eV for W = 4 [395]. The gap amplitude increases with the reduced W and relies on the ratio of the
removed Si atom and the total Si atom numbers of silicene [397]. Given a value of R, the opened band gap monotonically
decreases with the increasing even W. In the cases of R = 1 and 2, the maximum band gap is about 0.68 eV (Fig. 3d). The
mechanism of gap opening in SNM is the same as that of graphene nanomesh [393], that is, it has a geometric symmetry

Ó Macmillan Publishers Limited 2016

Fig. 80. (a) Geometric configuration of SNMs with R = 1 and W = 4. The yellow and white balls stand for Si and H atoms, respectively. Electronic band
structures of SNMs with odd (b) and even (c) W. (d) The band gap of SNM as a function of W with R = 1 and 2. (e) Band gap is plotted versus the quantity
N 1=2
removed =N total (Nremoved is the number of removed Si atoms, and Ntotal is the number of atoms before the holes is created in the unit cell). The black line
represents a linear fit to the data. Reproduced with permission from Ref. [397].
102 J. Zhao et al. / Progress in Materials Science 83 (2016) 24–151

Ó American Physical Society 2016


Fig. 81. Orbital-projected DOS and site-projected DOS for the FM defective SiNR with N1 = N2 = 4 symmetrically positioned extended defect. Reproduced
with permission from Ref. [401].

origin [394]. In [R, W = even] SNM, its K and K0 points of pristine silicene are folded into the U points of SNM. Degeneracy at
the Dirac points is lifted and a sizable band gap appears.
Similar to graphene nanomesh [393], the band gap Eg in [R, W = even] SNM is determined by the relation:

N1=2
Eg ¼ g removed
; ð23Þ
Ntotal
where Ntotal and Nremoved are the numbers of the total Si atoms before digging the holes and the removed hole atoms per unit
cell, respectively, and g is a fitting factor. The dependence of the band gap of SNM on N1/2 removed/Ntotal is shown in Fig. 80d,
yielding g = 7.246 eV. The perturbative model for silicene under periodic potential [397] can be used to understand the
reciprocal relationship between the band gap and Ntotal. The FET devices on the basis of silicene nanomeshes will be discussed
in Section 5.1.4.
Gong and co-workers [396] employed a combined DFT and TB approach to investigate the electronic band structure of
nanomeshes. In addition to the hexagonal Si6 antidot with an H-terminated edge, they also considered substitutional doping
of Al3N3 patch and its analogies, i.e., (ZnO)3, (AlP)3, (AlAs)3, (GaN)3, (GaP)3, and (GaAs)3, inside the silicene lattice. For both
hexagonal and orthogonal supercells with different sizes, they found that substitutional doping of Al3N3-type patches always
induces a band gap at the Fermi level while the opening of band gap by Si6 antidot depends on supercell size, which is
actually the distance between neighboring holes as found by the other groups.
In addition to the infinite systems discussed above, Aghaei and Calizo [398] considered the armchair silicene nanoribbons
perforated with periodic hexagonal holes. Their DFT calculations predicted that periodic nanoholes in ASiNRs can tailor the
electronic band gap in terms of the ribbon width, repeat periodicity, and position relative to the edges, opening a potential
route to silicene electronics.

4.6.4. Extended line defects


In the realistic samples of silicene sheets, co-existence of several domains is frequently observed. Therefore, 1D extended
defect like grain boundary in silicene should be inevitable. Till now, several kinds of line defects in silicene sheet have been
theoretically explored and some relevant effects on device applications have been discussed [384,400–404].
Due to the buckling nature, there are two equivalent energy-degenerate geometric phases in silicene: a phase has a given
atom shifted up and its first neighbors shifted down, whereas in b phase the shifts are reversed. Lima et al. [400] investigated
such kind of line defects in silicene created at the interface between a and b phases, which may occur along different
orientations (zigzag, armchair, or chiral geometries). The formation energies of these line defects are around 0.2 eV/nm,
which is one order of magnitude lower than the grain boundaries in graphene [405]. Interestingly, the defect-induced
electronic states appear far from the Fermi energy. Compared to the pristine region, these phase interfaces with enhanced
reactivity are preferential adsorption sites for foreign atom (like Au) or molecule, which may lead to more opportunities in
molecular engineering.
As for the topological line defects, Li et al. [384] considered a few possible configurations with different arranges of defect
rings: 5|7, 4|8, 5|8|5, d5d7, t5t7. Among them, 5|8|5 (shown in Fig. 77i) was found to be most energetically favorable due to
the mixed sp2/sp3 hybridization of silicon. Similar to graphene [405], all these extended 1D defects can significantly modify
the electronic properties of silicene by transforming semimetallic silicene into metallic.
Using AIMD simulations, Pati and co-workers investigated the structural reconstructions of different kinds of grain
boundaries in the silicene sheets [404]. They found that extended line defects (ELDs) such as 5–5–8, 4–8, 4–4–4, can be
incorporated into the boundary in a controlled way and the resulted structures are quite stable at rather high temperature
J. Zhao et al. / Progress in Materials Science 83 (2016) 24–151 103

(up to 1000 K) as well as on the Ag(1 1 1) surface at room temperature. Moreover, these ELDs can significantly modify the
electronic and magnetic properties of a freestanding silicene sheet, leading to new opportunities in silicene-based nanoelec-
tronic and spintronic devices.
Wood’s group [401] has further studied the spin-dependent electronic properties of 5|8|5 line defect in zigzag silicene
nanoribbons. As displayed in Fig. 81, it is possible to obtain half-metallic characteristic in defective SiNRs, where the majority
carriers are conductors and the minority carriers are semiconducting. The site-projected DOS in Fig. 81 demonstrates that
the majority DOS is composed mainly of the zigzag edge atoms (1) and the atoms on the defect line (8) and (9). The minority
DOS is much smaller and the peak at Fermi level is attributed to atoms (8), while the edge atoms (1) are responsible for the
two peaks in the valence and conduction bands.
Using a TB Hamiltonian including the SOC, Yang et al. [402] investigated the electronic structure of the 5|8|5 line defect in
silicene and found spin helical states around the line defect, which belong to electronic bands in the bulk energy gap. When
the system is subjected to random distribution of spin-flipping scatters, electron transport along the line defect suffers much
less spin-flipped scattering than in the bulk. Thus, an electric gate above the line defect can tune the spin-flipped transmis-
sion, implying application as a spin-controllable waveguide.
Very recently, Miwa et al. investigated the electronic properties of pristine and hydrogenated grain boundaries in silicene
by ab initio calculations [403]. Under an electric field perpendicular to the silicene layer, which removes the inversion
symmetry of silicene, valley-indexed metallic states ruled by the grain boundaries was found, characterizing the QVHE.
Moreover, hydrogen incorporation on the grain boundary would suppress those valley-indexed states due to a spin
polarization.

Fig. 82. (a) The atomic structure of silicene. The band structures of silicene with (b and c) the zero strain of e = 0, (d and f) the compressive strain of e = 0.08,
and (e and g) the tensile strain of e = 0.10. The Fermi level EF is indicated as the line at zero. Reproduced with permission from Ref. [408]. Copyright 2016,
Elsevier B.V.
104 J. Zhao et al. / Progress in Materials Science 83 (2016) 24–151

4.6.5. Adatoms
Si adatom on silicene lattice has also been theoretically considered by several groups [377,384,406]. As shown in Fig. 77g,
the Si adatom prefers the top site of silicene with the original lattice Si atom pressed down, forming a dumbbell configura-
tion perpendicular to the basal plane [377,406]. Owing to the satisfactory formation of local sp3 hybridization, the formation
energy for Si adatom is even slightly negative, i.e., 0.03 eV [377] or 0.06 eV [384]. Most interestingly, an individual Si
adatom not only opens a noticeable band gap (0.091 eV [377] at PBE level, see Fig. 79f; 0.5 eV at HSE06 level [406]), but also
induces a local magnetic moment of 2 lB in the silicene supercell [377,384,406]. In such way, an all-silicon magnetic semi-
conductor can be obtained. In addition to Si, first-principles calculations showed that adsorption of adatom of the other
group IV elements, i.e., C, Ge, Sn, would lead to similar effects on the structural and electronic properties of silicene [406].
Furthermore, the migration behavior of a Si adatom on silicene has also been simulated by DFT calculations. Two possible
diffusion paths of Si adatoms was proposed [377], both having energy barrier of about 1 eV, in contrast to the small diffusion
barrier of 0.3–0.4 eV for C atom on top of graphene [374]. Therefore, once adsorbed on silicene, the Si adatom is very like to
be immobilized due to the large migration barrier.

4.7. Strain effect

In the experimentally synthesized silicene sheets, certain strain may exist due to the lattice mismatch between the
silicene sheet and the substrate. On the other hand, change of electronic band structure under uniaxial or biaxial strain
provides new opportunities for modulation of the physical properties of silicene, rendering so called ‘‘strain engineering”.
So far, there have been a number of theoretical investigations on the strain effect of electronic structures of silicene using
first-principles calculations [69,163,407–410]. Zhao [69] presented DFT calculations on the electronic properties of silicene
under uniaxial tensile strain along AC and ZZ directions. Under small strain, a direct gap opened at K point increases with the
increasing of the strain, reaching a maximum value of 0.08 eV (0.04 eV) for ZZ (AC) oriented tension of 0.08 (0.05). After-
wards, the band gap decreases under large strain and vanishes before the low buckled silicene structure becomes unstable
at e = 0.14 (e = 0.18) for ZZ (AC) tension. Similar effect of gap opening in strained silicene was predicted by Mohan et al. [411]
in their first-principles calculations. They also discussed the strain effect on the optical properties in terms of dielectric func-
tion. Under tensile and asymmetric strains, the p plasmon in silicene disappears. Meanwhile, the p + r plasmon peak shows
small blue-shift under compressive strain and red-shift under (uni- and bi-axial) tensile strains, respectively, but does not
sensitively rely on asymmetric biaxial strain.
Using DFT calculations, Liu et al. [163] studied the effect of biaxial strain on the band structure of silicene. They found that
silicene retains its semimetallic characteristic up to a biaxial tensile strain of 7.5% and then becomes metallic under larger
strains. This semimetal–metal transition results from the downward shift of the lowest conduction band at the C point,
which is associated with weakening of the p⁄ bond. Similarly, Kaloni et al. [407] showed that the Dirac cone remains essen-
tially at the Fermi level in silicene under biaxial strain up to 5%. At higher strain, the Dirac cone shifts towards higher energy,
e.g., lying above the Fermi level by 0.06 eV at a strain of 7%. Therefore, strain can be used to induce hole doped Dirac states
in silicene, which is a consequence of weakened Si–Si bonds.
The strain-induced self-doping effect in silicene was further demonstrated by Wang and Ding [408], who considered both
tensile and compressive biaxial strains. As shown in Fig. 82, under the compressive strain, the Dirac point moves below the
Fermi level and renders n-type doping; on the contrary, under the tensile strain, the Dirac cone shifts upwards above the
Fermi level, resulting in p-type doping. This strain-induced self-doping effect is attractive for the future applications of
Ó American Physical Society 2016

Fig. 83. (a) Relative thermal conductivity of silicene sheet, silicene nanoribbons, graphene sheet, and graphene nanoribbons as a function of uniaxial strain
along the longitudinal direction. The reference values of j0 are 40.1, 31.4 and 35.0 W/mK for silicene and silicene nanoribbons (width of 5 and 10 unit cells),
and are 2002.2, 1323.3, and 1531.9 W/mK for graphene and graphene nanoribbons (width of 5 and 10 unit cells). (b) Percentage contribution to total heat
flux from vibrations in the x, y, and z directions as a function of uniaxial strain for silicene and graphene. Reproduced with permission from Ref.
[418].
J. Zhao et al. / Progress in Materials Science 83 (2016) 24–151 105

nanoelectronic devices for the following two reasons: (i) the shift of Dirac cone can be effectively tuned by the magnitude of
strain; (ii) silicene can be switched among behaviors of n-type, zero-gap, and p-type simply by applying strain. Note that
such effect only exists in silicene but does not occur in graphene, which can be understood by the fact that strain can alter
the atomic structure of the low-buckling silicene sheet significantly and thus affect the sp2/sp3 hybridized state.
The strain-induced shift of band structures also renders an effective way to tailor the work function of silicene, as demon-
strated by Qin and co-workers [412,413]. Up to a biaxial strain 15%, the WF of silicene increases monotonically from the
strain-free value of 4.79 eV to 5.12 eV, then it remains nearly invariant for larger strains. In contrast, the WF increases mono-
tonously with uniaxial strain in the entire elastic region for both AC and ZZ directions and does not saturate at large strains.
Up to a small uniaxial strain of 3%, the WF increases slightly by 0.01 eV regardless of the direction of strains. Then, the WF of
silicene under the ZZ strain increases more significantly than that under the AC strain. At the ultimate strains, i.e., 17% for AC
and 15% for ZZ, the work functions reach 4.98 and 5.11 eV, respectively.
It is worth to note that the previous ab initio results on whether a uniaxial strain open a gap in silicene is contradictory,
that is, Zhao [69] and Mohan et al. [411] predicted a gap opening effect, while the other groups [163,407,408,412,413] only
observed shift of energy bands. Recently, Voon et al. [414] derived an effective Hamiltonians for silicene with strain from
first-principles band structures and provided a formal resolution to the above controversy. They found that an in-plane
uniaxial strain leads to the same energy change for both bands making up the Dirac point since they both have the same
deformation potential; thus, the only effect should be a shift in band energies instead of gap opening. However, the influence
of strain does significantly change the dispersion at finite k values and consequently the Fermi velocity of those bands.
Indeed, first-principles calculations by Qin et al. [413] revealed anisotropic variation of Fermi velocity (up to about ±10%
under uniaxial strain of less than 12%) for different bands of silicene.
As the hydrogenated derivative of silicene, silicane is an indirect semiconductor with band gap of 2.94 eV (PBE) [415] or
4.15 eV (GW) [416]. The effect of tensile strains on the electronic structure of silicane has been discussed by several groups
[69,415,416]. At a small biaxial strain (1.6% [415] or 2.74% [416]) or a moderate uniaxial strain at AC direction (5% [69] or
6.3% [415]), the silicane becomes a direct-gap semiconductor. Regardless the loading condition (biaxial or uniaxial), the band
gap decreases monotonically with increasing strain, for instance, from 4.15 eV for pristine silicene to 2.58 eV for biaxial
strain of e = 10% [416]. Under sufficiently large strain (e.g., 20% AC strain and 30% ZZ strain [69]), the band gap would
be closed and silicane eventually becomes metallic. The lattice-scattering limited electron mobilities of pristine and strained
silicane have been computed by Restrepo et al. [415]. Although the strain-free silicane possesses a mobility of 464 cm2/V s,
which is about 1/3 of that of bulk silicon (1450 cm2/V s), the mobility of silicane increases significantly with increasing strain
after the indirect–direct transition, e.g., a high mobility to 8551 cm2/V s was achieved at a strain of 4%. As a high-mobility
direct-band gap semiconductor, the strained silicane would be an ideal candidate for future nanoelectronic devices like FET.
In addition to the electronic properties, Pei et al. [417] simulated the effect of uniaxial strain on thermal conductivity of
monolayer silicene sheet by using NEMD with MEAM potential. Regardless the orientation (AC or ZZ), the thermal conduc-
tivity of silicene first increases with increasing tensile strain from zero to e = 0.04, then it starts to decrease with further
increasing strain. At a tensile strain of 0.12, the thermal conductivity drops more than 30% compared to the strain-free value.
This behavior can be attributed to the unique low-buckled structure of silicene. At small tensile strains, the bucked config-
uration becomes less buckled due to bond rotation, resulting in the enhanced in-plane stiffness and thus an increase in the
thermal conductivity of silicene. At large tensile strains, the Si–Si bonds are severely stretched, which in turn reduces the
in-plane stiffness and the thermal conductivity of silicene.
Hu et al. also studied the thermal response of silicene sheet and nanoribbons of different widths subjected to uniaxial
stretching using NEMD method with Tersoff potential (Fig. 83) [418]. In contrast to graphene, the thermal conductivity of
silicene sheet and nanoribbons is found to increase dramatically with tensile strain, and reaches a plateau until the final
breakup. The narrower SiNR is more sensitive to strain than the wider SiNR. The phonon dispersions show that the uniaxial
strain softens the transverse and longitudinal acoustic and optical modes, and stiffens the flexural acoustic and optical
modes for silicene. The authors explained this abnormal thermal response of silicene to strain by considering the flexural
ZA and ZO modes as dominating the phonon transport in silicene [418], which however is under debate [72,75].
Besides the above mentioned NEMD simulations with empirical potentials, Xie et al. recently performed first-principles
calculations to predict the lattice thermal conductivity of silicene under biaxial strain by solving the phonon Boltzmann
Ó American Chemical Society 2016

Fig. 84. Atomic structures of silicene on: (a) h-BN; (b) Si-SiC; (c) C-SiC substrates. Reproduced with permission from Ref. [428].
106 J. Zhao et al. / Progress in Materials Science 83 (2016) 24–151

transport equation [419]. It was found that the thermal conductivity of silicene can increase dramatically with strain, which
was mainly attributed to the dramatic enhancement in the acoustic phonon lifetime due to the flattening of the buckled
silicene structure upon stretching.

4.8. Semiconducting and insulating substrate effect

For the real applications of silicene in nanoelectronic devices, silicene sheets have to be placed on semiconducting or
insulating substrate. Therefore, how semiconducting and insulating substrates affect the electronic properties of silicene
becomes an important issue and has attracted plenty of attentions so far. Houssa et al. proposed that a silicene flake could
preserve the semimetallic behavior when it is embedded between ultrathin insulating AlN layers [36]. SiC is frequently used
as an insulator in microelectronics industry and graphene-based FETs [420–422]. Meanwhile, h-BN monolayer was success-
fully fabricated in experiment [423] and widely utilized as substrate in graphene-based FET to improve the performance
[424–427]. Therefore, these two substrates have been frequently studied by first-principles calculations to explore the
substrate effect on silicene [428–430]. The structures of silicene on h-BN and SiC are shown in Fig. 84. For SiC, two kinds
of surfaces were considered, i.e. Si-terminated SiC(0 0 0 1) surface (Si-SiC) and C-terminated SiC(0 0 0 1) surface (C-SiC). It
was found that silicene preserves its low-buckled geometry with a distance of about 3 Å from all these substrates. The bind-
ing energies between silicene and h-BN and SiC substrates are about 0.07–0.09 eV per silicon atom, which correspond to the
typical strength of vdW interaction. Thus, these substrates can provide effective mechanical support for silicene without
disturbing its atomic structure.
The band structures of pristine silicene, silicene/BN, silicene/Si-SiC and silicene/C-SiC are presented in Fig. 85. The Dirac
cone in pristine silicene (highlighted by red lines) can be well preserved after being placed on h-BN and Si-SiC. Actually,
small band gaps of several meV are opened in silicene/BN and silicene/Si-SiC systems. For silicene supported on BN, several
Moiré superstructures have been studied and the band gap opened at the Fermi level reaches around 30 meV, which is nearly
independent of the rotation angle between the two lattices [429]. The intact linear dispersion of silicene is also confirmed
by the partial charge densities around the Fermi level. As shown in Fig. 85f and g, the partial charge densities in the energy
range of EF and EF 0.2 eV mainly distribute on silicene, indicating that the bands around Fermi level all come from silicene.
In sharp contrast, the p bands of silicene are severely disturbed by C-SiC substrate and silicene becomes n-doped (Fig. 85d).
The partial charge densities (Fig. 85h) clearly show that the valence bands of C-SiC already enter the vicinity of Fermi level of
the silicene/C-SiC hybrid system.

Ó American Chemical Society 2016

Fig. 85. Electronic band structures of (a) silicene, (b) silicene/BN, (c) silicene/Si-SiC, (d) silicene/C-SiC and partial charge densities of (e) silicene, (f) silicene/
BN, (g) silicene/Si-SiC, (h) silicene/C-SiC. The red lines highlight the p-bands of silicene in different configurations. The partial charge densities are in the
energy range between EF and EF 0.2 eV. The Fermi level is set to zero. The enlarged view of the band lines at K point near the Fermi level is presented as the
inset in (b and c). Reproduced with permission from Ref. [428].
J. Zhao et al. / Progress in Materials Science 83 (2016) 24–151 107

The different substrate effects on the electronic properties of silicene can be attributed to the relative locations of VBM
and CBM of silicene and substrates. The CBM (or VBM) of silicene is lower than the VBM of C-SiC substrate by 0.1 eV. Due to
the coupling between silicene and C-SiC substrate, some new electronic states in the energy range between the VBM of C-SiC
and silicene arise, lowering the total energy of the hybrid system. After the combination of silicene and C-SiC substrate, the
Fermi energy of the hybrid system will stay between the VBM of C-SiC and the VBM of silicene. Hence, the Dirac cone of the
pristine silicene is submerged under the Fermi level. While for the silicene/BN and silicene/Si-SiC systems, the CBM of
silicene locates in the gap region of BN and Si-SiC. The electronic states in the energy range between the VBM of Si-SiC
(or BN) and the CBM of silicene are forbidden. Consequently, the conical point of the Dirac cone in silicene is basically
retained at the Fermi level.
In a recent experiment [30], silicene has been successfully synthesized on a semiconducting substrate, MoS2 surface.
However, the prepared silicene is high-buckled and metallic, rather than the usual low-buckled silicene, as already discussed
in Section 3.2. Correspondingly, high-buckled silicene on MoS2 substrate have been theoretically investigated by DFT calcu-
lations [431–433]. According to the relative positions of MoS2 and silicene, six stacking patterns were constructed. The most
stable configuration for high-buckled silicene on MoS2 (HSMS), namely AA-I HSMS, is shown in Fig. 86a. The atomic struc-
tures for all the stacking patterns are similar, with a buckled height (d) of about 1.9 Å within silicene layer and an equilib-
rium distance (D) over 3 Å between silicene and MoS2 substrate. For all stacking patterns, high-buckled silicene can be
adsorbed on MoS2 substrate with binding energies around 80 meV per silicon atom, indicating a weak vdW interaction
between silicene and the substrate [432]. Due to the weak interaction, the metallic electronic structure of HB silicene is
preserved on MoS2 according to the calculated electronic band structure shown in Fig. 86b.
Regardless the experimental observation of high-buckled silicene, low-buckled silicene on MoS2 (LSMS) has also been
considered by first-principles calculations [430,432–434]. Abundant Moiré patterns exist for LSMS due to the large lattice
mismatch (17.9%) between silicene and MoS2 [432]. The energy differences between different Moiré patterns are very small
(several meV per silicon atom). One of the supercells of LSMS, namely (3 2/3 1)M LSMS, is shown in Fig. 86c, with d = 0.53 Å,
D = 3.10 Å and binding energy of 100 meV per silicon atom. The band structure of (3 2/3 1)M LSMS displayed in Fig. 86d
indicates that the Dirac point of silicene can be well preserved with a small band gap of 29.3 meV opened at the Dirac point.
Recently, a small gap of 40 meV at the Dirac cone was also predicted by DFT calculations with careful consideration of the
interlayer van der Waals interaction [433].
In addition to SiC, h-BN and MoS2, many other insulating or semiconducting substrates, such as MoX2 (X = Se, Te), WSe2,
silicane, Cl-silicene, Br-silicene, graphane, solid argon, MgX2 (X = Cl, Br and I), hydrogenated Si(1 1 1) surface, hydrogenated
Ge(1 1 1) surface, GaS, GaSe, GaTe and b-Si3N4(0 0 0 1)/Si(1 1 1), have been extensively explored as possible supports of
silicene [430,431,435–444]. First-principles calculations show that all these substrates can provide effective mechanical
support without destroying the geometry and electronic structures of silicene and result in small band gaps in the magnitude
of tens to one hundred meV. Among them, silicene on WSe2 exhibits structural uncertainty originated from the possibility of
lateral shifts between the silicene and WSe2 substrate, due to small translation energy barriers (less than 0.4 meV per silicon
Ó American Chemical Society 2016

Fig. 86. Geometries (a and c) and band structures (b and d) of AA-I high-buckled silicene on MoS2 (HSMS) and (3 2/3 1)M low-buckled silicene on MoS2
(LSMS). Reproduced with permission from Ref. [432].
108 J. Zhao et al. / Progress in Materials Science 83 (2016) 24–151

atom) [444]. Therefore, electronic structures of silicene on WSe2 can fluctuate between metallic and semiconducting even at
low temperature. On the other hand, silicene on S and Te doped WSe2 are metallic and semiconducting, respectively, because
of the large barrier for lateral shifts [444]. A band gap of 0.34 eV would be opened for silicene on Te doped WSe2 without
destroying the Dirac physics of pristine silicene.
In particular, by direct evaluation of the Z2 topological invariant as well as by using adiabatic continuity arguments,
Huang and co-workers proposed that the Bi-intercalated silicene/Si(1 1 1) system is a topologically trivial insulator [445].
Moreover, an out-of-plane external electric field turns the system into a Z2 topological insulator with a band gap larger than
that of freestanding silicene [445].
All these above mentioned substrates have inert surfaces without any dangling bonds and thus interact with silicene
through weak vdW interactions, which is responsible for the preservation of geometry and electronic structures of silicene.
However, if there are dangling bonds on the surface of substrates, the unique electronic properties of silicene would be
destroyed. For example, silicene loses its linear dispersion around Fermi level when it is placed on ZnS(0 0 0 1) surface and
the hybrid system was predicted to be semiconducting with an indirect gap of about 0.7 eV [446]. The electronic properties
of pristine silicene would be badly destroyed when epitaxially grown on the (1 1 1) surfaces of AlAs, AlP, GaAs, GaP, ZnS and
ZnSe [258]. Small band gap arises due to the symmetry breaking of two sublattices in silicene.
The dispersion relationship of silicene near Fermi level can be approximately described using a tight-binding model
[447]:
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
Eð~
2 2
kÞ ¼  D2 þ h m2F k ; ð24Þ

where k is the wavevector relative to Dirac point, mF is the Fermi velocity, and D represents the on-site energy difference of an
electron at the two sublattices. Accordingly, the band gap of silicene/substrate superstructures is Eg = 2D. Van der Waals
interactions yield small D and thus small band gaps. Inspired by this, the band gap created at the Dirac point may be further
tuned by the interaction strength between silicene and substrates or bias voltage. A recent study by Zhao’s group predicted
that the band gap increases with the decreasing interlayer distance between silicene and h-BN [429]. First-principles
calculations showed that the electronic properties (e.g., band gap and carrier effective mass) of the confined silicene between
diamond and silicon atomic layers can be widely tuned by changing interlayer spacing [448]. Moreover, the band gaps of
silicene/MoS2 [432,434], silicene/MgX2 (X = Cl, Br and I) [436], silicene/silicane [437] and silicene/GaS [439] hybrid systems
can be effectively modulated by external electric field perpendicular to the interface.
In addition to electronic properties, the substrate effect on the thermal transport properties of silicene has been investi-
gated by using NEMD simulations with Tersoff potential [84]. Silicene on amorphous SiO2 substrate manifests a thermal
conductivity of only about 20% of that of freestanding silicene at room temperature (Fig. 87a). Interestingly, the thermal
conductivity j shows a much weaker temperature dependence for the supported silicene (j  T0.5) relative to the
freestanding silicene (j  T1.5), indicating the dominating phonon-substrate scattering over phonon–phonon scattering
in presence of the substrate. The reduction of thermal conductivity in SiO2-supported silicene can be understood by the
mode-decomposed phonon relaxation time (Fig. 87b). When the silicene layer is placed on SiO2 substrate, the relaxation
time decreases significantly for the phonons of three acoustic branches, which contribute to most of the thermal conduction.
The interfacial thermal transport between silicene and various semiconducting or insulating substrates (i.e., crystalline
silicon, amorphous silicon, crystalline silica and amorphous silica) has been further investigated by NEMD simulation
[449]. It was found that the interfacial thermal resistances between silicene and all substrates decrease nearly 40% as Ó American Institute of Physics 2016

Fig. 87. (a) Thermal conductivity of freestanding silicene and silicene on SiO2 substrate along the armchair and zigzag directions as a function of
temperature. j0 = 48.1 W/mK is the thermal conductivity of freestanding silicene along the armchair direction at T = 300 K. The inset shows the
extrapolation of thermal conductivity at infinite sample length. (b) Relaxation time of different phonon branches for freestanding and supported silicene at
T = 300 K. Reproduced with permission from Ref. [84].
J. Zhao et al. / Progress in Materials Science 83 (2016) 24–151 109

temperature increases from 100 K to 400 K due to the enhanced phonon couplings and that amorphous substrates facilitate
the interfacial thermal transport compared with their crystalline counterparts.

4.9. Multilayer heterostructures

With the boom of 2D materials, multilayer heterostructures composed of different 2D atomic crystals become an attrac-
tive and important issue because that stacking different 2D atomic crystals can provide opportunities to create new layered
materials and thus bring in novel physical/chemical properties. The properties of multilayer heterostructures depend on the
composition of 2D crystals and the stacking style, and thus would be more tunable than those of single-component 2D mate-
rials. Up to now, graphene-based bilayer heterostructures, such as graphene/h-BN [450–452], graphene/MoS2 [452–456] and
graphene/WS2 [457], have been successfully fabricated in experiments and exhibit high performance as channels of FETs
[452,454,456,457]. Silicene, a new Dirac 2D material similar to graphene, is expected to play an important role in construct-
ing multilayer heterostructures. A number of theoretical studies on silicene multilayer heterostructure have been done in
recent years.
First, silicene/graphene bilayer heterostructure has been intensively investigated [458–462]. By making different combi-
nations of the supercells of silicene and graphene, various stacking patterns of silicene/graphene heterostructure with a
small amount of strain (0.5–2.7%) were constructed [459]. Silicene and graphene keep their original structures in all of these
p p p p
bilayer heterostructures and the interlayer distance is about 3.4 Å [458–460]. Taking ( 3  3) silicene on ( 7  7)
graphene [459] as an example, the binding energy between the two layers is 121 meV per Si atom. Note that the sliding bar-
rier is only 0.4 meV per Si atom, implying that silicene can easily slide on graphene due to weak vdW interaction. The band
p p p p
structure of ( 3  3) silicene/( 7  7) graphene heterostructure is depicted in Fig. 88. The projected band structures for
silicene and graphene as well as the combined system clearly demonstrate that the linear dispersions of silicene and
graphene are both well preserved in the bilayer heterostructure. Due to small amount of charge transfer from silicene to
graphene, silicene and graphene in the heterostructure is weakly p- and n-doped, with the Dirac cones shifting by about
0.1 eV above and below the Fermi level, respectively. Small band gaps are opened at the Dirac points, i.e., 26 meV for silicene
and 2 meV for graphene, also due to the weak interaction between two layers. For (2  2) silicene/(3  3) graphene hybrid,
the estimated Fermi velocities at the Dirac points are 0.5  106 m/s and 0.8  106 m/s for silicene and graphene [458],
respectively, indicating that the excellent electronic properties of both layers are preserved in the heterostructure. Interest-
ingly, the charge carrier concentrations of silicene and graphene decrease gradually with increasing the interlayer separation
and a conversion of doping types of silicene and graphene occurs when their interfacial distance artificially increases to
above 4.2 Å [458]. Therefore, changing the interlayer distance seems to be an effective way of tuning the electronic proper-
ties of bilayer heterostructure.
Recent DFT calculations combined with the nonequilibrium Green’s function formalism show that silicene/graphene
bilayer heterostructure exhibits enhanced transmission as compared to the sum of the transmissions of individual graphene
and silicene sheets at energies near the Fermi level [461]. However, away from the Fermi energy, some states become local-
ized in one of the layers, which reduce the electron transmission across the sample.
Besides electronic properties, the optical adsorption is enhanced in silicene/graphene heterostructure compared to the
silicene and graphene monolayers [458]. Moreover, the interfacial thermal conductance G of (2  2) silicene/(3  3)
graphene heterostructure was also studied using MD simulations [463]. The G of (2  2) silicene/(3  3) graphene
heterostructure at room temperature was calculated to be 11.74 MW/m2 K when the heat transfers from graphene to
silicene, and was 9.52 MW/m2 K from silicene to graphene, showing a thermal rectification ratio. A recent theoretical study
also proposed that graphene/silicene bilayer is a good candidate for a nanocapacitor with permanent dipole and piezoelectric
Ó American Physical Society 2016

p p p p
Fig. 88. Band structures of 3  3 silicene on 7  7 graphene: (a) the projected states on Si are highlighted; (b) the projected states on C are
highlighted; and (c) the projected bands in (a and b) are combined. The substrate-induced gap is about 26 meV for Si (C) and 2 meV for graphene (K),
respectively. Reproduced with permission from Ref. [459].
110 J. Zhao et al. / Progress in Materials Science 83 (2016) 24–151

capabilities [462]. With the outstanding electronic, optical and thermal properties all together, silicene/graphene
heterostructures show great potentials in future applications of nanoscale devices.
Beyond the silicene/graphene bilayer heterostructures, a graphene/silicene/graphene three-layer heterostructure was
explored by Neek-Amal and co-workers using first-principles calculations [259]. In this three-layer heterostructure, silicene
with a buckled height of 0.41 Å is placed just in the middle of two planar graphene layers with the interlayer distance of
3.42 Å, similar to that in silicene/graphene bilayer heterostructure. The linear dispersion around the Dirac cones of silicene
and graphene are well preserved without any band gap opened at the Dirac cones, which is different from that in silicene/
graphene bilayer heterostructure. This is attributed to the symmetric structure of the three-layer heterostructure, in which
no symmetry breaking of two sublattices in silicene occurs. However, silicene and graphene are p- and n-doped, respectively,
which is the same as that in bilayer heterostructure. Using reactive molecular dynamics and DFTB simulations, Berdiyorov
et al. proposed that the three-layer heterostructure may be realized by intercalating Si atoms in between bilayer graphene
[464]. Moreover, a superlattice made by alternate stacking of graphene and silicene shows similar electronic properties to
that of silicene/graphene heterostructure within the in-plane direction [465]. However, the calculated band structure
indicates that the superlattice is metallic in the direction perpendicular to the graphene layer.
A sandwich hybrid of silicene/BN/silicene, whose atomic structure is illustrated in Fig. 89a, was proposed by Kamal and
co-workers [466]. Band structures of this silicene/BN/silicene heterostructure at different interlayer distances are displayed
in Fig. 89. One can see characteristic linear dispersion around EF at its equilibrium configuration (Fig. 89b) due to weak
interlayer interaction. However, the nature of dispersion curve becomes parabola-like when the system is compressed along

Fig. 89. (a) Atomic structure and (b–f) band structures of silicene/BN/silicene heterostructure at different interlayer distances. The interlayer distance is
given in the plot of each band structure in unit of Å. Reproduced with permission from Ref. [466]. Copyright 2016, Elsevier B.V.
Ó American Chemical Society 2016

Fig. 90. (a) Atomic structure of BN/silicene/BN heterostructure. (b) Band gap as a function of vertical electric field E\ for BN/silicene/BN heterostructure and
silicene. Reproduced with permission from Ref. [31].
J. Zhao et al. / Progress in Materials Science 83 (2016) 24–151 111

Fig. 91. Band gap (Eg) as a function of vertical electric field (E) for various silicene based heterostructures: (a) silicene/MoS2. Reproduced with permission
from Ref. [434]. Copyright 2016, Royal Society of Chemistry. (b) Silicene/GaS. Reproduced with permission from Ref. [439]. Copyright 2016, American
Institute of Physics. (c) Silicene/silicane. Reproduced with permission from Ref. [437]. Copyright 2016, Elsevier B.V. (d) Silicene/DL-H. Reproduced with
permission from Ref. [437]. Copyright 2016, Elsevier B.V.

the direction perpendicular to the basal plane (Fig. 89c). Finally, opening of a band gap near the Fermi level occurs when the
inter-layer distance is about 3 Å and below (Fig. 89d–f).
Similarly, a superlattice of silicene and h-BN was constructed by alternatively stacking these two kinds of layers [467].
The resulted superlattice retains the Dirac cone of freestanding silicene at the Fermi level and thus exhibits semimetal
character. However, the system undergoes an interesting phase transition from a semimetal to a topological insulator and
further to a band insulator under an external electric field perpendicular to the silicene layers with increasing field strength.
Another sandwich structure, BN/silicene/BN (Fig. 90a), is also a zero-gap semiconductor with Dirac cone at the Fermi level
[31]. A small band gap can be opened when a vertical electric field (E\) is applied. Interestingly, the band gap increases
linearly with the strength of electric field and the band gap of sandwiched heterostructure is larger by 50% than that of
freestanding silicene under the same E\ (Fig. 90b).
In addition to the heterostructures discussed above, some semiconductor heterostructures base on silicene were also
proposed by theoretical calculations, including silicene/monolayer MoS2 [434], silicene/monolayer MoSe2 [468], silicene/
monolayer ZnS [469], silicene/GaS nanosheet [439], silicene/ultrathin silicon nanosheet [437] and silicene/silicane [470].
Monolayer MoS2, monolayer MoSe2, monolayer ZnS, GaS nanosheet are all calculated to be semiconductor with band gap
of 1.84 eV [434], 1.4 eV [468], 2.56 eV [469] and 2.6 eV [439], respectively. Several kinds of ultrathin silicon nanosheets were
considered, including one layer silicon atoms saturated by hydrogen atoms (HSiH or silicane), fluorine atoms (FSiF), one-side
hydrogen atoms one-side fluorine atoms (HSiF) as well as two-layer silicon atoms saturated by hydrogen atoms (DL-H),
which are all semiconductors [437]. The band structures of these heterostructures are basically the simple superposition
of the two parts of the heterostructure due to the weak vdW interaction between the compositing layers
[434,437,439,468,469]. Thus, electronic bands near the Fermi level mainly come from silicene. All these heterostructures
are semiconductors since that small band gaps are opened at the Dirac point due to the symmetry breaking of two
sublattices in silicene [434,437,439,468,469].
A common feature of these heterostructures is that their band gaps can be effectively tuned by a vertical electric field.
However, the details for the variations of their band gaps with electric field are different. For silicene/MoS2, under a positive
external electric field, Eg monotonously rises up to 191 meV (Fig. 91a). When a negative external electric field is applied, Eg
first decreases and disappears under E = 0.5 V/Å, then it progressively increases up to 75 meV under E = 1.0 V/Å (Fig. 91a).
The amplitude of Eg is nearly linearly dependent on E with a slope of 0.15 eÅ [434]. The effect of electric field on the band gap
can be understood by the following picture. The interaction between silicene and MoS2 introduces on-site energy difference
of two sublattices in silicene and thus opens a band gap at zero electric field. A vertical electric field can also result in on-site
energy difference of the two sublattices in silicene due to the buckled configuration of silicene, which is confirmed by the
tunable band gap of freestanding silicene under vertical electric field [31,32]. Therefore, the external electric field can either
enhance or weaken the on-site energy difference depending on its direction. This picture is suitable for all heterostructures
112 J. Zhao et al. / Progress in Materials Science 83 (2016) 24–151

discussed here. For silicene/GaS (Fig. 91b), the system suddenly becomes metal at E = 0.7 V/Å, because of the shift of bands
at the C point [439]. For silicene/silicane (Fig. 91c), a step appears at 0 < E < 0.01 V/Å, in which the band gap varies slightly
with the electric field. Note that the band gap changes from direct gap at K point to indirect gap at C point when
E > 0.13 V/Å, which results in a turning point at E = 0.13 V/Å (Fig. 91c) [437]. For silicene/DL-H, a step similar to that in
silicene/silicane emerges in the electric field range of 0.016 < E < 0.01 V/Å, in which the band gap changes slightly with
the field [437].
Apart from external electric field, strain can also modulate the band gaps in silicene/ZnS, silicene/GaS and silicene/DL-H
heterostructures [437,439,469]; while interlayer distance and strain can also effectively modulate the band gap of the
silicene/ZnS and silicene/silicane heterostructures [469,470]. Different from the silicene/MoS2 heterostructure, the superlat-
tice composed of alternative stacking of silicene and MoS2 layers exhibits metallic character [471].
In a recent theoretical study using DFT calculations, silicene encapsulated between two MX2 (M = Mo, W; X = S, Se, Te)
layers was predicted as a novel 2D topological insulator with a robust nontrivial band gap suitable for room-temperature
applications, which has significant implications for novel QSHE devices [472].

4.10. Substitutional doping and hybrid sheet

Substitutional doping is a common approach in condensed matter physics and materials science to tailor the electronic
properties of a material. Although it remains a big challenge to substitutionally dope silicene nanoribbons or sheets in exper-
iment, there have been numerous theoretical predictions on the doped silicene nanoribbons and nanosheets, showing very
promising results on the modulation of electronic and magnetic properties.

4.10.1. Substitutional doped silicene nanoribbon


The substitutional doping of an individual B or N atom as well as B/B, N/N or B/N pair in silicene nanoribbons has been
investigated by several groups [473–478]. It was found that single B or N dopant atom prefers to stay on the edge of nanorib-
bons. For armchair NR, incorporation of an individual B or N atom on one side would transform it from semiconductor into
metal due to the half-filled impurity-induced band. The ASiNR doped with two B or N atoms on two opposite edges is also
metallic. In contrast, ASiNR doped with a pair of B/N atoms is semiconductor with smaller band gap than the pristine form,
because of the charge compensation between N and B atoms. For zigzag NR, incorporation of one N or B impurity results in
asymmetrical spin-up and spin-down bands, transforming the antiferromagnetic nanoribbon into ferromagnetic. ZSiNR
doped with one N atom remains semiconducting, while that doped with one B atom exhibits half-metallic character. Under
a transverse electric field of 0.15 V/Å, single N doped zigzag nanoribbon becomes a half-metal. Doping a pair of B/B or N/N
atoms on the edge of ZSiNR would suppress the spin polarization at both Si edges and make the system non-magnetic. More
interestingly, doping a B/N pair at the edge or subedge of zigzag NR renders a half-metal and spin gapless semiconductor
with 100% spin polarized currents around the Fermi level.
Similar to the above works on B or N doping, silicene nanoribbons substitutionally doped with Al or P have been theo-
retically explored [473,477,479–481]. Chen and Ni [479] investigated hexagonal AlSi armchair nanoribbons with different
widths, which can be regarded as doped silicene NRs with injected holes. They found these hole-doped NRs show itinerant
magnetism originated from delocalized and dispersionless energy bands near the Fermi level. Zhang et al. [481] studied the
doping effect of single P atom in silicene NRs with both armchair and zigzag edges. The doped P atom prefers to stay on the
edge of ASiNR and the system becomes ferromagnetic semiconductor due to the impurity-induced spin splitting. Meanwhile,
P-doped ZSiNR is ferromagnetic, similar to the case of N doping. When a ZSiNR is doped with single Al atom at certain sites, it
exhibits half-metal behavior with 100% spin polarization, which is desirable for spintronics [477].
In addition to band structures, electron transport properties of zigzag silicene nanoribbons doped with B, N, Al, P atoms
have been computed [480,482]. It was found that the presence of a single Al or P impurity induces quasibound states and
thus new dips in the electron conductance of nanoribbon [480]. When doped in the central region, the Al (P) atom acts as
an acceptor (donor), just like the case of conventional semiconductor. However, this behavior is reversed if the dopant is
located on the edge sites. Similar results were observed in the case of two (Al or P) impurities, but the conductance was
modified by the impurity–impurity interaction. Lopez-Bezanilla [482] further demonstrated that an individual impurity
introduces asymmetric electron–hole conductivities that depend on the type and position of dopant atom. Randomly
distributed impurities over 1 lm long and 4 nm wide silicon nanoribbons widen the intrinsic electronic gap due to quantum
confinement. The mobility gaps created at low doping rates may be useful for the design of new silicene devices.
In addition to B, N, Al, P dopants, doping of single or double carbon chains in zigzag silicene nanoribbons have been
considered by Song et al. [483–485]. They found that the carbon atoms form a straight line instead of a zigzag one upon
relaxation, along with transverse contradiction of ribbon width. Zigzag silicene NR doped with single carbon chain is metallic
and ferromagnetic, while the detailed electronic and magnetic properties can be modulated by the position of carbon chain.
On the other hand, zigzag silicene NR decorated with double carbon chains on the two opposite edges are metallic and
antiferromagnetic.

4.10.2. Substitutional doped and hereto silicene sheets


The optical properties such as electron energy loss spectra (EELS), dielectric function, and optical absorption coefficient of
freestanding silicene (modeled by a 32-atom supercell) with various concentrations of Al, P and Al/P pair dopants have been
J. Zhao et al. / Progress in Materials Science 83 (2016) 24–151 113

simulated using DFT by Das and co-workers [486]. For all doped systems, the maximum values of the absorption coefficients
are higher than the pristine one, and the reflectivity modulation is restricted to low energy (<4 eV) and high energy (>8 eV)
for the parallel and perpendicular polarizations, respectively. After P doping, a new additional EELS peak is observed in
perpendicular polarization. A sudden jump of optical conductivity emerges for the Al–P co-doped system at 18.75% doping
concentration. These theoretical predictions may be useful guidance for fabricating silicene-based optoelectronics devices.
Moreover, a recent first-principles study suggested that germanium doping in silicene can significantly reduce its thermal
conductivity by 62% for a doping concentration of 6% without affecting the electronic transport properties [487]. The depres-
sion of phonon transport is attributed to the low-frequency phonon softening and enhanced phonon scattering by Ge doping.
As extremes cases of substitutional doping with very high dopant concentration, a series of silicene based heterosheets
with honeycomb lattice can be obtained. Some of them are shown in Fig. 92 as representatives. For instance, using DFT and
CALYPSO program, Tang et al. [488] predicted that the global minimum of BSi3 monolayer is a silicene-like planar sheet
containing aromatic Si6 rings (Fig. 92a). Such abnormal planar geometry was attributed to the strong p–p conjugation
between Si6 and B atoms. This 2D BSi3 silicene and its various 1D derivatives (nanotubes and nanoribbons) are all metallic,
and the metallic behavior is very robust to mechanical strain and surface chemical functionlization. Therefore, these BSi3
nanomaterials are expected to serve as anodes in high-capacity Li ion batteries and other energy-storage devices.
Using first-principles calculations, Ding and Wang [489] systematically explored a series of XSi and XSi3 sheets (X = B, C,
N, Al, P) with honeycomb lattice. The Si-based heterosheets are all mechanically stable and possesses versatile buckled
structures and electronic properties, as summarized in Table 6. The equilibrium geometries of these hybrid sheets can be

Fig. 92. Atomic structures of various silicene-based heterosheets with honeycomb lattice: (a) BSi3; (b) SiB; (c) SiC; (d) SiP. Note that (a and c) are flat, while
(b and d) are buckled.

Table 6
Structural, energetic and electronic information for the SiX and XSi3 (X = B, C, N, Al, P) hetero sheets [489].

Buckling type Dh (Å) Ecoh (eV/atom) Eform (eV/atom) Band gap (eV)
SiB Low washboard 0.35/0.46 4.55 0.32 Metal
SiC Flat 0 5.96 0.023 2.54
SiN High washboard 1.22/0.19 5.07 0.51 Metal
SiAl Low washboard 0.42/0.48 3.10 0.15 Metal
SiP High chair 0.94 3.64 0.001 Metal
BSi3 Flat 0 4.27 0.12 Metal
CSi3 Flat 0 4.77 0.16 Semimetal
NSi3 High chair 0.82/0.59 4.40 0.16 Metal
AlSi3 Flat 0 3.59 0.15 0.95
PSi3 High chair 0.60/1.16 3.78 0.001 0.57
114 J. Zhao et al. / Progress in Materials Science 83 (2016) 24–151

flat (SiC, BSi3, CSi3, AlSi3), chair-like buckling (SiP, NSi3, PSi3), or washboard-like buckling (SiB, SiN, SiAl). The corresponding
electronic properties of these hybrid sheets are diverse, including metals (SiB, SiN, SiAl, SiP, BSi3, NSi3), zero-band gap
semimetal (CSi3), narrow-band gap semiconductors (AlSi3, PSi3), and semiconductor (SiC).
Among the silicene-based monolayer hereto sheets, considerable attentions have been paid to the silicene–graphene
[490–494] hybrid sheets. Different from the low-buckled silicene, the most stable form of SiC is planar hexagonal lattice
and possess a direct band gap of about 2.5 eV (GGA) [491] or 3.48 eV (GW) [493]. As the ratio of Si atoms in the Si-C hybrid
sheet increases from 50% to 100%, an increase of buckling and a narrowing of band gap were simultaneously observed [492].
Although the most stable hexagonal SiC (h-SiC) monolayer is semiconductor, some metastable conformers of SiC monolayer
composed of Si–Si and C–C pairs, i.e., t1-SiC and t2-SiC, are semimetal and possess Dirac cones due to a unique pair coupling
mechanism [494]. From the GGA calculation, full hydrogenation of the h-SiC monolayer would widen its band gap to 4.02 eV,
while half-hydrogenation on the every Si (C) atoms would lead to a ferromagnetic semiconductor with smaller band gap of
0.81 eV (0.34 eV) [490].
There have been also some studies on the silicene–germanene hybrid sheets [495–497]. Zhou and co-workers [496] found
that SiGe monolayer inherits the low-buckled honeycomb structure of pristine silicene with slightly larger buckling height
(0.579 Å) and is mechanically stable without imaginary frequency in the phonon dispersion curve. Naturally, Dirac cone state
and high Fermi velocity were observed in the electronic band structures of SiGe sheet, similar to silicene. Moreover,
half-hydrogenation on either Si or Ge atoms leads to a ferromagnetic semiconductor (1 lB per H), whereas hydrogenation
of every Si atom is energetically preferred and the resulting HSiGe monolayer sheets possesses a sizeable band gap of
1.495 eV from HSE06 calculation. Monte Carlo simulations with a 2D Ising model predicted that the Curie temperature of
this ferromagnetic HSiGe monolayer sheet is about 110 K. Similarly, Zhang et al. [497] suggested a ferromagnetic
half-metal with 1 lB per formula unit by bromination on every Si atoms of the SiGe monolayer, which exhibits an even
higher Curie temperature of 240 K.
In addition to SiGe monolayer, using DFT calculations and a special quasi-random structure model, Padilha et al. [495]
showed that the 2D SixGe1x random alloys in the honeycomb lattice for any composition x are nontrivial topological
insulators with the same topological classification Z2 = 1 as pristine silicene and germanene. The band gap can be further
tuned by an external electric field, while the systems keep the nontrivial topological characteristics up to a critical value
of |Ec| = 1 V/nm.
Besides doping with hetero elements, isotope doping is an effective way to reduce the thermal conductivity of a system
without affecting its electronic properties. The isotope dopants alter the phonon dispersions and act as scattering centers to
cause phonon localization. The room-temperature phonon transport properties of silicene nanosheets doped with 29Si, 30Si
and 42Si isotopes have been investigated based on NEMD method using Tersoff and SW potentials (Fig. 93a) [78]. For silicene
with random isotope doping, the thermal conductivity attains minimum at an isotope concentration of 50%, and a reduction
of about 20% is achieved for 30Si doping. The thermal conductivity can be even lowered by alternative patterning 28Si and 30Si
roads of the same width perpendicular to the phonon transport direction (Fig. 93b). The interfaces between the 28Si and 30Si
patches scatter phonons, and the thermal conductivity reaches a minimum at a periodic length of 2.76 nm, corresponding to
a reduction of 30% relative to that of pristine 28Si silicene. Roughing the 28Si/30Si interfaces leads to additional deterioration
to the thermal conduction, and the thermal conductivity is reduced to about 58% of that of 28Si silicene.

5. Silicene electronics and spintronics

5.1. Simulation of silicene FETs

Opening a sizable band gap (>0.4 eV) without degrading the electronic properties is critical for the application of silicene
in nanoelectronics. Although covalent functionalization or patterning in the form of nanoribbon can open a sizable band gap
Ó IOP Publishing 2016

Fig. 93. (a) Thermal conductivity of silicene nanosheets as a function of doping percentage. (b) Thermal conductivity as a function of temperature for
pristine and isotope doped silicene nanosheets. Reproduced with permission from Ref. [78].
J. Zhao et al. / Progress in Materials Science 83 (2016) 24–151 115

in silicene, the electronic properties are destructed, and the carrier mobility is inevitably reduced to a large extent, as
previously found for graphene.
Up to now, there are two mechanisms to open a band gap in silicene while preserving its electronic properties: one is the
inversion symmetry breaking and the other is intervalley interaction. More specifically, applying a vertical electric field to
silicene layer is a typical way to open a band gap by breaking the inversion symmetry between the two sublattices [31];
creating a silicene nanomesh [397] by creating periodic holes in silicene is an effective way of opening a band gap without
destructing the electronic properties by intervalley interaction, which has been discussed in Section 4.6.3. The surface
adsorption of foreign atom [33,34] (as summarized in Section 4.4) is a way to open a band gap in silicene based on a mixture
of the two mechanisms. With the opened band gap, various kinds of silicene FETs can be theoretically designed and their
performances will be discussed below in details.

5.1.1. Effects of vertical electric field


Applying a vertical external electric field E\ will give rise to a difference in the electrostatic potentials of the two sublat-
tices in monolayer silicene because of its buckling structure. As a result, the inversion symmetry between the two sublattices
is broken, and hence a direct band gap Eg is opened at the K and K0 point in the Brillouin zone. According to two independent
DFT calculations [31,32], the opened band gap increases linearly with E\ as shown in Fig. 94d. This linear dependence of Eg
on E\ can be well explained in terms of a TB model, which yields Eg = eE\D (D is the buckling height in a silicene layer)
[31,32,498]. The band structures of silicene around the Dirac cone at three different E\ are displayed in Fig. 94a–c. Typically,
the band gap is Eg = 0.08 eV under E\ = 0.51 V/Å. The rate at which the band gap opens is 0.157 eÅ, greater than the result of
0.0742 eÅ obtained by Drummond et al. [32]. Possible reasons for this discrepancy include differences in the size of vacuum
and basis set during computations. The effective masses of electrons in silicene monolayer at E\ = 0.4 V/Å are mKe C = 0.015m0
and mMKe = 0.033m0 (m0 is the free electron mass), which are comparable with those in bilayer graphene (mKe C = 0.028m0 and
MK
me = 0.012m0) without electric field. The mobility of suspended silicene under E\ = 0.4 V/Å is estimated to be in the order of
105 cm2/(Vs) by assuming the relaxation time is the same for the two systems using the relation l = es/m⁄.

5.1.2. Dual gated silicene FET


A dual-gated silicene FET with SiO2 dielectric and h-BN buffer layer was theoretically investigated by Lu’s group, as
illustrated in Fig. 95a [31]. The doping level as well as the vertical electric field in this device can be controlled by the
dual-gate. The vertical electric field applied to the silicene is E\ = (Vt  Vb)/(d0  2di + 2di/e), where Vt and Vb are top and
bottom gates, respectively. The total gate voltage Vg = Vt + Vb reflects the total doping level. When E\ = 0, a transmission
‘‘bulge” around EF occurs inside the bias window at Vg = 0.2 V (Fig. 95b). However, under external field E\ = 1 V/Å, an
obvious transport gap of about 0.14 eV is observed. Under Vg = 1 V, the transport gap moves outside of the bias window,

Ó American Chemical Society 2016

Fig. 94. (a–c) Band structures of silicene around EF at three different vertical electric fields. Inset in (a) band structures in the first Brillouin zone at E\ = 0.
The Fermi level or the valence band top is set to zero. (d) Dependence of band gap and charge transfer per Si atom on the electric field E\. Reproduced with
permission from Ref. [31].
116 J. Zhao et al. / Progress in Materials Science 83 (2016) 24–151

(a)

(b) (c)
0.3
Vg = -0.2 V, E⊥= 0 Vg = -0.2 V, E⊥= 0
6
Vg = -0.2 V, E⊥= 1 V/Å Vg = -0.2 V, E⊥= 1 V/Å

Ó American Chemical Society 2016


Vg = 1 V, E⊥= 1 V/Å Vg = 1 V, E⊥= 1 V/Å
0.2
Transmission

PDOS
0.1
2

0.0 0
-0.50 -0.25 0.00 0.25 0.50 -0.50 -0.25 0.00 0.25 0.50
Energy (eV) Energy (eV)

(d) Vg = -0.2V

Vg = 1V
Phase (radians)

0 π 2π

Fig. 95. Dual-gated silicene FET: (a) schematic model, (b) transmission spectra, and (c) projected density of states of the channel under different gate
voltages and electric fields with a fixed Vbias = 0.1 V, obtained from the GGA/SZ level. The dashed vertical line indicates the bias window. The channel length
is 67 Å. The Fermi level is set to zero. (d) Transmission eigenstates for the off- and on-state (Vg = 0.2 and 1 V, respectively) at Ef and at the (0, 1/3) point of
k-space under E\ = 1 V/Å. The isovalues are 0.8 a.u. Reproduced with permission from Ref. [31].
Ó Macmillan Publishers Limited 2016

Fig. 96. Schematic model of the FET with the Na-adsorbed silicene as channel. The channel is 113.8 Å long, and the electrodes are composed of semi-infinite
silicene. Reproduced with permission from Ref. [34].

and the device is turned from off-state to on-state. The on/off current ratio under E\ = 1 V/Å is 4.2 at 300 K. Such low on/off
current ratio may be attributed to the short channel (67 Å) that leads to a large tunneling leakage current in the off-state.
The sp3 hybridized silicene is reactive. On the common dielectric layer of SiO2, silicene would covalently bond with Si and
O atoms. A true dielectric material that will not disturb the electronic properties of silicene is thus desirable. As discussed in
Sections 4.8 and 4.9, silicene layer is approximately intact when placed on or sandwiched between h-BN layers even under
an electric field of up to 2 V/Å. Therefore, it was strongly recommended that insertion of an h-BN buffer layer between
J. Zhao et al. / Progress in Materials Science 83 (2016) 24–151 117

silicene and oxide substrate in a silicene FET device would be beneficial to maintain the structural integrity and high carrier
mobility.
Beside h-BN layer, other substrates including hydrogenated Si-terminated SiC(0 0 0 1) surface, graphane, and MoS2, also
interact with silicene via weakly vdW forces [428–430,434]. Therefore, silicene on these substrates can retain its geometry
structure and the excellent electronic properties according to ab initio calculations [466,467], as discussed in Section 4.8. The
insertion of alkali atoms between silicene and metal substrate can also restore the Dirac cone of silicene with n-type doping
character [196]. The Dirac cone of silicene can also be preserved when embedded between two graphene layers [259]. How-
ever, the hybrid structure is metallic since both silicene and graphene are semimetallic and the minima of their projected
density of states do not coincide with each other.

5.1.3. Alkali-adsorbed silicene FET


As discussed in Section 4.4, adsorption of alkali metal atoms can induce band gap up to 0.5 eV in silicene. Moreover, a
large on/off current ratio in this structure based FET is expected [34]. An FET composed of Na adsorbed silicene channel
(50% coverage), and semi-infinite silicene electrodes has been simulated, as shown in Fig. 96. The dielectric region consists
of a SiO2 substrate and h-BN buffer layers to preserve the high carrier mobility of silicene. Below the dielectric region, a
bottom gate is placed.
When Vg = 0 V, the FET is in the on-state: the transport gap is located below EF and a broad transmission peak locates
within the bias window (Fig. 97a). As Vg decreases, EF shifts downward (right insets of Fig. 97a); Vg decreases to 18 V, EF
moves closer to the transport gap. When Vg = 30 V, EF is located in the gap region and the transmission probability nearly
vanishes within the transmission window. Consequently, an effective off-state with a tiny current of 5.45  107 lA/lm is
reached. The transmission eigenstate at EF and k point = (0, 1/3) in Fig. 97b also reflects the switch effect: the transmission
eigenvalue in the on-state is 0.70, and the corresponding incoming wave function reaches the other lead with slight scatter-
ing. In the off-state, the transmission eigenvalue is only 0.01, and the incoming wave function cannot reach the other lead
due to strong scattering. The FET is n-type, as shown in the transport characteristics in Fig. 97c. The on/off current ratio in
this FET is very high (4  108) as expected.

5.1.4. Silicene nanomesh FET


While metal atom adsorption can induce a large band gap up to 0.5 eV in silicene [33,34], a drawback of this method is the
requirement of a large supply voltage (Vdd) of 30 V for FET applications. Designing a new silicene FET with a high on/off
ratio under a low supply voltage is highly desired. As discussed in Section 4.6.3, SNM has a maximum band gap of 6.8 eV,
quite suitable for FET applications. In Fig. 98a, the schematic model of a single-gated FET based on the SNM with semi-
infinite silicene electrodes is presented. As shown in Fig. 98b, the transfer characteristics of the sub-10 nm SNM FETs demon-
strate distinct on/off current modulation. The on/off ratio in the 9.1 nm SNM FET can reach 5.1  104 (Vdd = Von  Voff = 0.5 V
and Vg = 0.5 V is chosen as the on-state), which already meets the requirement of 104–107 for the high-speed switches. The
subthreshold swing (SS = dVgate/d(log I)) is 68 mV/dec, approaching the 60 mV/dec switching limit of the classical transistors.
The on/off current ratio Ion/Ioff decreases greatly from 5.1  104 to 17 as Lgate decreases from Lgate = 9.1 nm to 3.8 nm, showing
a significant short-channel effect (the gate voltage window is limited to a supply voltage of 0.5 V).
Ó Macmillan Publishers Limited 2016

Fig. 97. Device performance of the Na-covered silicene FET. The bias voltage is set to 0.1 V. (a) Transmission spectra with Vg = 0, 18 and 30 V,
respectively. The vertical dashed-line indicates the bias window. The left insets show the details in the bias window. The right insets illustrate the relative
positions of Ef and the gap. (b) Transmission eigenstates (at E = EF and k point = (0, 1/3)) of the on-state (Vg = 0 V) and off-state (Vg = 30 V). The isovalue is
0.2 a.u. (c) Transfer characteristic in linear (left axis) and logarithmic scales (right axis). Reproduced with permission from Ref. [34].
118 J. Zhao et al. / Progress in Materials Science 83 (2016) 24–151

Ó Macmillan Publishers Limited 2016


Fig. 98. (a) Schematic of a SNM FET structure. The yellow and white balls stand for Si and H atoms, respectively. (b) Transfer characteristics of the SNM FETs
for different channel lengths at Vbias = 0.2 V. Reproduced with permission from Ref. [397].

Table 7 compares the critical performance parameters of the sub-10 nm SNM (9.1 and 7.8 nm), advance Si, and CNT FETs
at Vbias = 0.5 V. The gate control ability of dual gated FET is obviously better than single gated FET. The 9.1 nm dual-gated
SNM has a better performance than the sub-10 nm advanced Si devices and 9 nm CNT device. While the 7.8 nm
dual-gated SNM has a better performance than the sub-10 nm advanced Si devices and is competitive with the 9 nm CNT
device. Although the 9.1 and 7.8 nm dual-gated SNM FETs both delivers a high on-state current, their on/off current ratios
are only 1.8  103 and 1.2  103, respectively, which cannot meet the requirement of HP logic of ITRS.
After considering the phonon effect (only contribution by elastic scattering), the on-state current and the on/off current
ratio at Vbias = 0.5 V in a 9.1 nm single-gated SNM FET decrease by 1–2 orders. Therefore, the sub-10 nm SNM FETs has an
excellent performance at zero temperature, which is, however, greatly degraded at room temperature when phonon scatter-
ing effect is included.

5.1.5. Tunneling FET


As discussed above in Section 5.1.3, ultrahigh on/off current ratio can be obtained in the alkali metal adsorbed silicene
FET, but the required supply voltage Vdd is large. The power dissipation has been a fundamental issue for nanoelectronics.
Tunneling FETs (TFETs), based on band-to-band tunneling instead of thermal fluctuation, can have a small subthreshold
swing below the limit of 60 mV/dec in MOSFET and supply voltage Vdd, allowing for less power dissipation.
Simulation of a silicene p–i–n TFET composed by three types of doping in different regions of silicene (Fig. 99) shows an
on-state current of 1000 lA/lm, SS of 77 mV/dec, and supply voltage of 1.7 V [33]. The greatly enhanced on-state current
and reduced supply voltage suggest decreased power dissipation.

Table 7
Comparison of performance metrics between sub-10 nm SNM, advance Si, and CNT transistors [397].

Channel Lch (nm) Vbias (V) Vdd = Von  Voff (V) Ion (lA/lm) Ion/Ioff SS (mV/dec)
SNM (single-gated) 7.8 0.5 0.5 205 3.3  103 72
SNM (dual-gated) 7.8 0.5 0.5 607 8.9  103 68
0.71 0.71 1963 1.2  103 74
SNM (single-gated) 9.1 0.5 0.5 464 7.4  103 82
SNM (dual-gated) 9.1 0.5 0.5 870 1.2  104 74
0.72 0.72 3122 1.8  103 113
Si nanowire [499] 10 0.5 0.5 300 1.0  104 89 (Vbias = 1.0 V)
Si Fin [500] 10 0.5 0.5 138 1.0  103 125 (Vbias = 1.2 V)
ETSOI [501] 8.0 0.5 0.5 41 1.0  104 83 (Vbias = 1.2 V)
CNT [502] 9.0 0.5 0.5 630 1.0  104 94
HP logica 8.9 0.72 0.72 1350 1.35  104
HP logicb 8.0 0.71 0.71 1330 1.33  104
a
HP logic technology requirements of ITRS in 2022.
b
HP logic technology requirements of ITRS in 2023.
J. Zhao et al. / Progress in Materials Science 83 (2016) 24–151 119

Chemistry 2016
Ó Royal Society of
Fig. 99. Schematic model of the silicene TFET. The yellow silicon atoms form a sheet of monolayer silicene. The source and drain are Ir (blue atoms)/Cu (red
atoms) doped silicene and thus are p+/n+ type semiconductors with band gaps of about 0.3 and 0.2 eV, respectively. The central region, or the channel, is
doped with Pt (black atoms) and thus is a neutral type semiconductor with a gap of 0.3 eV. The concentration of the dopants in all parts of the device is
fixed at 5.6% (TMSi18, TM = Cu, Pt, and Ir). A 0.5 nm thick HfO2 dielectric region is placed over the channel. A thin h-BN buffer layer is utilized to protect
silicene from the oxide. Reproduced with permission from Ref. [33].

5.1.6. All-metallic FET


The low on/off current ratio due to the zero band gap limits the real applications of pure Dirac materials in nanoelectron-
ics at room temperature. Besides searching for effective approaches to open the band gap in silicene, all-metallic FET might
be another solution for silicene nanoelectronics. Without opening the band gap, the vertically stacked silicene/graphene as a
Dirac material can be directly used in high-performance devices [460]. The on/off switch effect comes from the forbidden of
electron transport from graphene to silicene near EF without assistance of phonon due to momentum mismatch and vice
versa, as shown in Fig. 100a.
The ab initio quantum transport simulation of the graphene/silicene heterostructure demonstrates a large transport gap of
over 0.4 eV, and the on/off current ratio is as high as 106 (Fig. 100d). The excellent performance of this all-metallic device is
robust against the relative rotation of the two Dirac materials and can be extended to silicene/germanene and homogenous
twisted bilayer graphene.

5.1.7. Silicene nanoribbon FET


Electron transport simulation of the armchair-edged silicene nanoribbons (ASiNRs) FETs (Fig. 101) shows tunable prop-
erties by changing width and length of the nanoribbon channels [503]. The band gap of silicene nanoribbon and thus the on/
off current ratio of silicene nanoribbon FET depend on the ribbon width. The on/off current ratio also positively relates to the
length of the channel in the checked range (44–115.1 Å). The calculated highest on/off current ratio in the examined devices
is 106. The subthreshold swing monotonously decreases with increasing L, and the lowest value calculated is 90 mV/dec.
The bottom of Fig. 101 depicts the output characteristic of the ASiNR (L = 9.89 nm) FET. At the above-threshold region
(Vgate P Vth  1 V), the current first increases linearly (Vbias < 0.1 V) and then becomes a constant (Vbias P 0.1 V) with the
increasing bias voltage. The output characteristics of many carbon-based (e.g., graphene [420,504], graphene nanoribbons
[505], and carbon nanotubes [506]) FETs exhibit either a linear shape without any saturation or only weak saturation.

Fig. 100. Graphene/silicene heterostructure and vertical FET: (a) Band structure. Green and red stand for graphene and silicene components. (b) (E, kx)
dependent transmission probability. Inset: Dirac cones of graphene (green) and silicene (red) in the miBZ. (c) Schematic of the single-gated heterostructure
vertical FET. (d) Transfer characteristics at Vds = 0.2 V. Reproduced with permission from Ref. [460]. Copyright 1999–2016, John Wiley & Sons, Inc.
120 J. Zhao et al. / Progress in Materials Science 83 (2016) 24–151

Ó Springer 2016
Fig. 101. Top: A gated two-probe model constructed by an optimized 6-ASiNR connected to the 4-ZSiNR electrodes. The channel length is 4.62 nm. Yellow
ball: Si; white ball: H. Bottom left: calculated transfer characteristics (Vbias = 0.02 V) of the 6-ASiNR (L = 9.89 nm) FETs. Bottom right: calculated output
characteristic of the 6-ASiNR (L = 9.89 nm) FET. The gate voltage varies from 0.5 to 3.0 V in a step of 0.5 V. Reproduced with permission from Ref.
[503].

The saturated output characteristic of the ASiNR-based FETs is an advantage to carbon based FETs. Up to now, both the
output current saturated [507] and unsaturated characters have been obtained [508,509] in the Si nanowire based FETs.

5.2. Experimental progress in silicene FETs

As discussed above, excellent and promising performances have been predicted in silicene FETs; but it remains very
challenging to realize these devices in experiment. The main obstacles are the difficulty of isolating silicene from its template
and the exposed silicene is unstable in air; thus the widely used wet transfer technique cannot be used to transfer silicene.
Recently, the key role of Si–Ag interaction in stabilizing silicene has been clearly pointed out by Tao et al. [42]. Using a synth
esis–transfer–fabrication process as shown in Fig. 102a, they successfully fabricated the first conceptual silicene FET. The

Ó Macmillan Publishers Limited 2016

Fig. 102. (a) Schematics of the synthesis–transfer–fabrication process of silicene. (b) Drain current Id versus gate voltage Vg curve displays ambipolar
electron–hole symmetry expected from silicene. (c) R versus gate overdrive voltage (Vg  Vdirac) of the silicene device. Measured transfer characteristics
(dots) are in good agreement with a widely used ambipolar diffusive transport model (line). Reproduced with permission from Ref. [42].
J. Zhao et al. / Progress in Materials Science 83 (2016) 24–151 121

first step is to epitaxially grow silicene on Ag(1 1 1), and then a 5 nm-thick layer of alumina is added above silicene. Second,
the silicene, Ag, and alumina layers are peeled together and flipped so as that Ag is on the top, and then put on a SiO2 sub-
strate. Therefore, the silicene layer is preserved by the attached Ag during transferring. Finally, part of Ag is etched away, and
the exposed silicene is the channel region and the left Ag film in the two ends acts as source/drain.
The expected Dirac-like ambipolar charge transport character was observed in this device (Fig. 102b and c). The current
on/off ratio is around 10 and the estimated band gap is 210 meV. The measured carrier mobility is 100 cm2 V1 s1, one
order of magnitude less than the predicted intrinsic mobility of silicene [76] (see Section 2.6). This may originate from
the acoustic phonon-limited transport and grain boundary scattering due to the substrate effect.
The key parameters for the performance of silicene based FETs are summarized in Table 8. Theoretically, a band gap with-
out degrading its high carrier mobility in silicene can be opened, and high-performance is expected in silicene transistors.
The experimentally achieved silicene FET shows an Ion/Ioff of 10 with the band gap of 0.21 eV. By taking full advantage of
the proposed approaches of band gap opening in Section 5.1, there is still large room for improving the performance of
silicene FETs.

5.3. Silicene spintronics

5.3.1. Spintronics in silicene nanoribbons


As we discussed in Section 4.5, silicene nanoribbons with different kinds of edges bring in new opportunities in silicene
research due to the rich possibilities of magnetic states. As a result, a variety of spintronic devices have been proposed based
on SiNRs. In the following, we will briefly discuss the magnetism and half metallicity of SiNRs and then give an overview of
the relevant spintronic applications.
Due to the edge magnetic states coupling, the antiferromagnetic (AFM) state of ZSiNRs is slightly lower in energy than the
ferromagnetical (FM) and nonmagnetic (NM) states [366]. However, because of the spin degeneracy, the AFM state cannot be
utilized directly in the spin-polarized transport (Fig. 103). Various approaches were proposed to break the spin degeneracy in
ZSiNRs, such as applying external electric field [366] and edge modification [362,368,510], just like in the case of zigzag
graphene NRs [511–513].
The AFM state of ZSiNRs possesses peculiar spin orbital distributions [366]. The zero-field a-spin and b-spin orbitals in
the conduction and valence bands are localized at the opposite edges of nanoribbon, and at the same edge the spin orien-
tations in the conduction and valence bands are opposite (Fig. 104c). By applying an in-plane homogenous electric field Eext,

Table 8
Performances of the simulated and experimental silicene transistors.

FET devices Channel length (Å) Band gap (meV) Ion/Ioff SS (mV/dec)
Theory Dual gated silicene FET [31] 67 160 4.2 –
Na-adsorbed silicene FET [34] 113.8 500 4  108 140
Silicene TFET [33] 32 200–300 1000 77
Silicene nanomesh FET [397] 91 680 7.4  103 (100a) 82
Silicene nanoribbon FET [503] 44–115.1 320–440 1–106 60–2616
All metallic silicene FET [460] 100 400b 106 1000
Experiment Silicene FET [42] 18,000 210 10 –
a
Phonon scattering effect included.
b
Transport gap.
Ó Royal Society of Chemistry 2016

Fig. 103. Band structures of (a) silicene, (b) 5-ZSiNR with AFM state, and (c) 5-ZSiNR with FM state. Reproduced with permission from Ref.
[514].
122 J. Zhao et al. / Progress in Materials Science 83 (2016) 24–151

Ó World Scientific Publishing Company 2016


Fig. 104. Electronic properties of the 6-ZSiNR in the ground state. (a) Spatial spin density distribution. The isovalue is 0.006 a.u. (b) Spin-resolved band
structures under Eext = 0, 0.1, and 0.25 V/Å, respectively. Inset: the band structure with Eext = 0.25 V/Å in the range of |E| < 0.1 eV and 0.7p/a 6 k 6 p/a (the
horizontal line is EF). The valence top or EF is set to zero. (c) a-spin and b-spin orbitals of the conduction and valence bands, shown as the square of the
absolute value of the wavefunction summed over all k-points. The isovalue is 0.275 a.u. The yellow arrow represents the energy shift direction of the spin
states under a transverse electrical filed. Blue and red denote a-spin and b-spin, respectively. Reproduced with permission from Ref.
[366].

Ó American Institute of Physics 2016

Fig. 105. The spin-polarized band structures for the asymmetric 6-ZSiNRs. Reproduced with permission from Ref. [362].

the half-metallicity can be induced in ZSiNRs. As Eext increases, the spin degeneracy of the conduction and valence bands is
lifted. The band gap of b-spin state decreases and finally closes under Eext = 0.25 V/Å, while that of a-spin state increases
slightly relative to the zero-field value. Therefore, half-metallicity in ZSiNRs is well established (Fig. 104b).
As discussed in Section 4.5, asymmetric edge modification is another way to transform ZSiNRs into half-metals
[362,368,510]. Taking asymmetric edge hydrogenation (H2–ZSiNR–H) as an example [362,368], several theoretical works
have demonstrated that the ground state of H2–ZSiNR–H is FM semiconductor. Around the Fermi level, the states are
completely spin-polarized with opposite spin orientations, declaring a bipolar magnetic feature in the H2–ZSiNR–H
(Fig. 105) [362]. Therefore, after appropriate doping, the Fermi level would shift into the valence or conduction bands,
and then H2–ZSiNR–H becomes a half metal.
J. Zhao et al. / Progress in Materials Science 83 (2016) 24–151 123

Spin filter – Based on the half-metallicity realized by the above methods, spin-filters can be modeled and simulated using
DFT coupled with nonequilibrium Green’s function (NEGF) formalism. A ZSiNR spin-filter under a transverse electric field is
illustrated in Fig. 106a [366]. In the spin-resolved transmission spectrum of this device under Egate = 0.3 V/Å, there is a large
peak for a-spin around EF and a clear gap for b-spin (Fig. 106b). Consistently, the spin-resolved transmission eigenstates at EF
and the C point in k-space in Fig. 106c also reflect the highly spin polarization. The spin-filter efficiency can be defined as:
SFE = (Ia  Ib)/(Ia + Ib), where Ia and Ib denote the a-spin and b-spin current densities, respectively. The calculated T(EF) and
SFE as a function of Egate are shown in Fig. 106d. The T(Ef) for the two spins are symmetric about Egate = 0. Even at small
Egate = 0.05 or 0.05 V/Å, T(Ef) between the two spins has a difference with a SFE = 62.3%. The difference becomes more
and more significant with the increasing Egate, consistent with the electric field-induced change in the band gap of ZSiNRs.
SFE is nearly saturated (99%) for electric field strength from |Egate| > 0.2 V/Å. Therefore, the dual-gated finite ZSiNR can serve
as a nearly perfect spin filter, with sign switchable by altering the electric field direction. A perfect spin filtering effect can
also be expected in H2–ZSiNR–H devices. The studies show that the SFEs of H2–ZSiNR–H devices with antiparallel (AP)
magnetic configuration can reach nearly 100% for |Vb| > 0.2 V as shown in Fig. 107 [368].
Spin FET – An effective spin FET model based on ZSiNRs was also proposed by adding a quadruple-gate, as illustrated in
Fig. 108 [366]. The electric field of the left pair of electrodes (ELG) is fixed and the on–off switch is achieved through
modulating the right one (ERG). When ERG = ELG, the left and right parts of the nanoribbon allow the same spin to transport
along the edges, corresponding to the on-state. By shifting the direction of the electric field of the right pair of electrodes, the
sign of the allowed travelling spin in the right part of the nanoribbon is reversed and contrary to that in the left part,

Ó World Scientific Publishing Company 2016

Fig. 106. Spin-filter based on the 4-ZSiNR. (a) Schematic model with one pair of gate electrodes on the two sides. (b) Spin-resolved transport spectrum
under Egate = 0.3 V/Å. (c) Spin-resolved transmission eigenstate at Ef and the C point in k-space under Egate = 0.3 V/Å. The isovalue is 1.0 a.u. (d) Spin-resolved
transmission coefficient at EF and spin filtration efficiency as a function of Egate. Reproduced with permission from Ref. [366].

Fig. 107. Spin-filtering efficiency for the H2–5ZSiNR–H and H2–6ZSiNR–H devices in the (a) negative and (b) positive bias, respectively. Reproduced with
permission from Ref. [368]. Copyright 2016, Elsevier B.V.
124 J. Zhao et al. / Progress in Materials Science 83 (2016) 24–151

Publishing Company 2016


Ó World Scientific
Fig. 108. Schematic model of the quadruple-gated spin FET based on the 4-ZSiNR. Reproduced with permission from Ref. [366].

resulting in a blockade of the transmission for both spins. As a consequence, the current of this device is expected to be for-
bidden in this case, which corresponds to the off-state. Therefore, through altering ERG, the quadruple-gated device can oper-
ate as a spin FET.
Giant magnetoresistance – Giant magnetoresistance (GMR) has been predicted in zigzag graphene NRs [515]. Inspired by
this work, a magnetoresistive device was proposed by making use of the electronic properties of ZSiNRs [515]. The AFM
ground state of ZSiNRs is semiconducting. Therefore, when connecting two metallic pristine silicene electrodes, as schema-
tized in Fig. 109a, it acts as a tunnel barrier that suppresses the conductance. The FM state of ZSiNRs is metallic. By applying a
proper magnetic field, ZSiNRs can switch between the semiconducting AFM and the metallic FM configurations
(Fig. 109a and b), and consequently a large current difference can be achieved (Fig. 109c). A large magnetoresistance (MR)
is thus realized.
Usually the optimistic definition is adopted to calculate MR, which is defined as MR(Vbias) = (IFM  IAFM)/IAFM, where Vbias is
the applied bias voltage, IFM (IAFM) is the total current of the FM (AFM) configuration. As illustrated in Fig. 109d, the MRs of

Ó Royal Society of Chemistry 2016

Fig. 109. Schematic models of a H-passivated ZSiNR connected to two semi-infinite silicene structures: (a) 3-ZSiNR in the AFM configuration, (b) 4-ZSiNR in
the FM configuration. Applying or removing magnetic field adjusts the ferromagnetic coupling of the two edges. The yellow (gray) balls denote silicon
(hydrogen) atoms. The arrows represent spin directions on the edges. (c) I–Vbias characteristics of the 5-ZSiNR at the AFM and FM configurations. (d) MRs of
different-width ZSiNRs as a function of bias voltage. Reproduced with permission from Ref. [514].
J. Zhao et al. / Progress in Materials Science 83 (2016) 24–151 125

different ZSiNRs at the same bias voltage drop generally with increasing ribbon width, except for 4-ZSiNR. In the examined
bias range from 0.05 to 0.5 V, the maximum MRs of the ZSiNRs are 261–1960%.
The critical magnetic field B⁄ required to switch a ZSiNR between the AFM and FM configurations can be estimated from
the following relation:
DE
B ¼ ; ð25Þ
g lB MT

where DE is the energy difference between the AFM and FM configurations, g = 2 is the Landauer factor of silicene,
lB ¼ 0:058 meV/T is the Bohr magneton, and MT is the total spin on edge atoms. The resulting B⁄ of the 5-ZSiNR is 8.6 T,
which is attainable in the laboratory.
Peculiar symmetry-dependent transport properties of FM ZSiNRs can be also utilized to achieve a GMR [264]. Due to the
existence of a twofold axis, the p and p⁄ wavefunctions in even-N ZSiNRs have opposite parity with respect to the C2
operation. The transmission of electron from the p band of the left electrode to the p⁄ band of the right electrode is forbidden,
leading to a conductance gap near the Fermi level. On the other hand, the transmission is allowed in odd-N ZSiNRs because
that the p and p⁄ wavefunctions have no definite parity.
Two spin configurations (P and AP) are considered in a two-probe system of even-N ZSiNRs. Both the left and right
electrodes are up-spin polarized in parallel (P) configuration (Fig. 110b), while in antiparallel (AP) configuration the two
electrodes have antiparallel spin polarization direction (Fig. 110c). For P configuration, the spin-up p bands and the spin-
down p⁄ bands of the two electrodes overlap near the Fermi level, and the transmission is allowed, as shown in Fig. 110f.
Therefore, near the Fermi level, the conductance is about 1 G0 for both spin components, and the current increases with
increasing bias. The situation is different in AP configuration. As illustrated in Fig. 110g, near the Fermi level, the spin-up
p⁄ (spin-down p) band of left electrode only overlaps with the spin-up p (spin-down p⁄) band of the right electrode. Because
the p and p⁄ bands have opposite parity, the transmission between them is forbidden. As a result, a conductance gap appears
around the Fermi level for both spin components, and the corresponding current is almost zero. Fig. 110d gives the I–V rela-
tionship for different spin components. The magnitude of MR can be evaluated according to the definition: MR = (IP  IAP)/IAP,
where IP and IAP are currents in P and AP configurations, respectively. The spin-up, spin-down, and total MRs are all in the
order of 104 (Fig. 110e).
SOC effect – All the above results about ZSiNRs are based on the DFT or DFT-NEGF calculations without considering SOC.
By using the Kane–Mele–Hubbard Hamiltonian model, the interplay of spin–orbit interactions and magnetic order have been
studied in general zigzag nanoribbon systems with honeycomb structure, which certainly includes ZSiNR [516]. The
Kane–Mele–Hubbard Hamiltonian gives:
X X X
H¼ tcþir cjr þ it SO rv ij cþir cjr þ U ni" ni# ; ð26Þ
hiji;r hhijii;r i

where r ¼ 1 is the projections of spin along the axis perpendicular to the nanoribbon, the single angled brackets stand for
first neighbor and double angled brackets for second, mij ¼ 1 for clockwise or anticlockwise second-neighbor hopping, and
in the third term Hubbard interaction ni" ¼ cþ i" c i" denotes the occupation operator of site i with spin up along an arbitrary
quantization axis. The Hubbard interaction is treated in the collinear mean field approximation, enforcing the magnetization
!
to lie along the quantization axis X ¼ ðsin a; 0; cos aÞ (Fig. 111a).
As shown in Fig. 112, for both FM and AFM edge magnetizations, the off-plane magnetization (a = 0) leads to a conducting
behavior, while the in-plane magnetization (a = p/2) opens a gap. As shown in Fig. 111b, the total energy E(a) reaches a min-
imum at a = p/2, i.e., for the in-plane magnetization, which means that the ground state of this system exhibits insulating
behavior. The band gap as a function of angle a is provided in Fig. 111c, that is maximal for in-plane magnetization
(a = p/2) and null for off plane. The gap opening occurs as long as magnetization is not off plane. Compared with the prop-
erties derived from the non-SOC DFT calculations, the electronic properties change dramatically after including the SOC,
which will certainly has dramatic consequences on the transport properties along the edges of ZSiNRs. However, all the
above discussed methods are still valuable guidance for device development. For instance, by applying a proper magnetic
field, the switch from the insulating in-plane magnetization ground state to conducting off-plane magnetization state can
also be achieved, thus realizing an effective MR.

5.3.2. Spintronics in semihydrogenated functionalized silicene


As we have discussed in Section 4.1, semihydrogenated silicene (denoted as H@Silicene thereafter), which is shown in
Fig. 113a and b, is dynamically stable as demonstrated by molecular dynamics and phonon dispersion calculations [517].
The ground state of H@Silicene is a FM semiconductor with a band gap of 0.93 eV (Fig. 113c). Regarding the spin-
polarized band structures, the Fermi level is shifted into the valence or conduction bands by applying an out-of-plane gate
voltage. Consequently, a half metal behavior is realized in the H@Silicene.
After the realization of half-metallicity, silicene-based spin filters and spin switches can be designed. A single-gated
H@Silicene device model as shown in Fig. 114a is simulated by using the DFT-NEGF method. As displayed in Fig. 114b, when
Vg = 0 V, the transmission probability nearly vanishes, which is responsible for the off-state for spin up and spin down. As Vg
increases, EF shifts upward (Fig. 114c). When Vg = 1.9 V, EF is located in the middle of a subband of spin-down and an effective
126 J. Zhao et al. / Progress in Materials Science 83 (2016) 24–151

Ó American Institute of Physics 2016


Fig. 110. (a) Band structure of FM 6-ZSiNR. The spin-up and spin-down components are presented in red and blue, respectively. (b and c) Spin density for 6-
ZSiNR with P and AP configurations under zero bias. The two electrodes have parallel spin configuration in the P case but antiparallel spin configuration in
the AP case. Pink and blue surfaces denote the spin-up and spin-down components, respectively. The isosurface corresponds to 0.01 e/Å3. (d) The I–V curves
for the P and AP configurations. The inset is semi-logarithmic scale plot. (e) The spin-up, spin-down, and total magnetoresistance on semi-logarithmic scale.
(f and g) The band structures for left and right electrodes, and the transmission spectrum for P (f) and AP (g) configurations under zero bias. Solid and open
circles denote p and p⁄ states. Solid arrows indicate allowed transmissions, and dashed arrows indicate forbidden transmissions. G0 equals to e2/h.
Reproduced with permission from Ref. [264].

Ó American Physical Society 2016

!
Fig. 111. Evolution of electronic properties for the FM ribbon as a function of the magnetization direction X ¼ ðsin a; 0; cos aÞ. (a) Scheme of the edge
magnetization for two different angles: 0 and a. (b) Total energy (per unit cell, with two magnetic atoms per cell), (c) gap, and (d) magnetization as a
function of a. Reproduced with permission from Ref. [516].

on-state for spin-down is achieved (Fig. 114d). With the increase of the gate voltage, the current density of the spin-down
channel increases significantly, but that of the spin-up value increases slightly, as shown in Fig. 114e. As a result, the SFE
increases with the increasing gate voltage and reaches 100% at Vg = 1.9 V.

5.3.3. Silicene spintronic applications


Magnetism could be introduced in silicene by the magnetic proximity effect with a magnetic insulator EuO. In a normal/
ferromagnetic/normal silicene junction, the charge, valley, and spin conductance oscillate with the length of the ferromag-
netic silicene [518,519]. The current through this junction is valley- and spin-polarized due to the coupling between valley
and spin degrees of freedom. By applying a local gate voltage, valley- and spin-polarizations can be tuned and a fully valley-
and spin-polarized current can be finally achieved.
A spin filter can also be designed based on a quantum point contact (QPC) in a silicene sheet characterized by a short and
narrow constriction to utilizing, as illustrated in Fig. 115a [128]. To make the valley degrees of freedom well separated, the
J. Zhao et al. / Progress in Materials Science 83 (2016) 24–151 127

Ó American Physical Society 2016

Fig. 112. Four different ferromagnetic configurations, either in- or off-plane and parallel (FM) or antiparallel (AF) edge magnetization, together with their
band structures calculated within the mean field Kane–Mele–Hubbard model. (a and b) Off-plane and conducting (both for FM and AF arrangements)
magnetizations. (c and d) In-plane and insulating parallel (both FM and AF) magnetizations. Calculations are done with U = t and tSO = 0.02t. Reproduced
with permission from Ref. [516].

zigzag edges are adopted for the whole geometry along the direction of current flow. The two opposite wide regions are in
the spin-valley-polarization metal (SVPM) phase with EF > 0 to model metallic source and drain (Fig. 115c) and the gated
constriction is in the M-VPM phase with a fixed Ez > Ec and an applied Zeeman field M ¼ lEz  kSO . An electrostatic potential
barrier U(xi) is added along the current flow direction, which is non-vanishing only in the constriction region. The resulting
spin polarization as a function of the effective chemical potential, l0 = EF  U0, is displayed in Fig. 115b. Through quantum
transport calculations based on the iterative Green’s function method, the QPC produces an almost fully spin-polarized cur-
rent. The spin-polarization direction can be easily reversed by locally changing the potential barrier via gating control in the
constriction. Such a high-efficiency, field-tunable spin filter based on silicene takes advantage of charge carriers in the bulk
system, small Rashba SOC, and controllable spin-splitting due to inversion symmetry breaking, and is robust against weak
disorder (compared with kSO þ M) and edge imperfections.
A Y-shaped spin/valley separator based on silicene (Fig. 116) was proposed to separate the two spin/valley polarizations
from the incoming lead 1, with one flowing to lead 2 and the other flowing to lead 3. An out-of-plane electric field, Ez > Ec, is
first applied in the central silicene sheet to create a nonvanishing Berry curvature, and chemical potential l is tuned into the
128 J. Zhao et al. / Progress in Materials Science 83 (2016) 24–151

Fig. 113. The optimized atomic configurations and the structural parameters of H@Silicene. (a) Top view with the rhombus marked in black shows the
supercell. The Bravais lattice vectors of unit cell are given with a1 = a2 = a = 3.899 Å. (b) The side view. The average bond length d1 (Å) between Si and H
atoms, d2 (Å) between Si and Si atoms, and buckled height DZ (Å) between Si and Si layers. The yellow and white balls stand for Si and H atoms, respectively.
(c) Band structures of H@Silicene. The arrow denotes the spin-polarized direction. The top of the valence band is set to zero. Reproduced with permission
from Ref. [517]. Copyright 2015, Elsevier B.V.

conduction bands. Then by setting potentials, for example, V1 > V2 = V3, at the terminals of silicene, charge carriers acquire an
anomalous velocity proportional to the Berry curvature in the transverse direction, similar to that reported by Xiao et al. in
graphene [520]. By linear response theory with negligible kR1 and kR2 , the spin and valley Hall conductivities at the Fermi
level in the conduction band are obtained as:
 
e2 kSO
rxy ðspinÞ ¼  ; ð27Þ
h l
 
e2 lEz
rxy ðvalleyÞ ¼ 1 : ð28Þ
h l
Valley and spin polarization imbalance at output terminals V2 and V3 are then obtained (with opposite polarization
between them). Therefore, silicene provides an ideal platform for efficiently manipulating spin/valley degrees of freedom.

6. Other applications of silicene

6.1. Thermoelectric devices

The thermoelectric effect converts heat to electrical energy and vice versa. Extensive efforts have been devoted to explor-
ing novel thermoelectric materials, aiming for development pf nanoscale devices as power generators or cooling systems
[521–523]. Owing to the excellent scalability and compatibility with current silicon-based technology, integration of silicene
into nanoscale thermoelectric devices is tempting and currently of great interest.
The efficiency of a thermoelectric device is characterized by its thermoelectric figure of merit ZT defined as:

ZT ¼ S2 rT=ðje þ jp Þ; ð29Þ

where S is the Seebeck coefficient (thermopower), r is the electrical conductivity, T is the absolute temperature, and je and
jp are the thermal conductivities contributed by electrons and phonons, respectively. Typically, a materials with ZT  1 is
regarded as good thermoelectric material, while a device with ZT > 3 is competitive to the conventional energy conversion
J. Zhao et al. / Progress in Materials Science 83 (2016) 24–151 129

Fig. 114. (a) Schematic of two-probe model of H@Silicene sheet spin filter device with SiO2 dielectric and h-BN buffer layer. (b, c and d) Spin-polarized
transmission spectra with Vg = 0, 0.9 and 1.9 V, respectively. The bias voltage is fixed at Vbias = 0.2 V. Red (blue) line stands for the spin-down (up). The
vertical dashed-lines denote the bias voltage window. The insets are the schematic of the Fermi level shift with the gate voltage. (e) Spin filter efficiency and
spin-resolved current as a function of the gate voltage. Reproduced with permission from Ref. [517]. Copyright 2015, Elsevier B.V.

techniques [524]. Improvement of ZT can be achieved by either increasing the power factor S2r, or decreasing the total ther-
mal conductivity je + jp. ZT can also be expressed as:
ZT ¼ ZTe =ð1 þ jp =je Þ; ð30Þ
where ZTe = S2rT/je is defined as the electronic figure of merit, and it provides an upper limit of ZT [524].
Silicene possesses similar electronic properties as that of graphene, i.e., the characteristic Dirac cone and high carry
mobility, as we discussed in Section 2. However, the lattice thermal conductivity of silicene is much lower than that of
graphene, due to its buckling structure [72,75]. Therefore, silicene and its nanostructures show great potentials for thermo-
electric devices. Intensive works have been done to investigate the thermoelectric effect in silicene sheets and nanoribbons
[524–529], which will be discussed in the following.
Due to the gapless feature, silicene has a small Seebeck coefficient of about 87 lV/K at room temperature [530], close to
that of graphene (100 lV/K) [338]. The maximum of the electronic figure of merit ZTe is 0.36, achieved at chemical poten-
tials of about ±0.08 eV, corresponding to electron (+) and hole () doping, respectively [524]. Placing silicene on metallic
substrates induces distortions of the atomic structure and hence alters the electronic bands. Silicene on Ag(1 1 1) exhibits
a buckling distance of 0.79 Å, rather than 0.43 Å in freestanding silicene, opening a band gap of about 0.3 eV (Fig. 117b).
130 J. Zhao et al. / Progress in Materials Science 83 (2016) 24–151

Ó Macmillan Publishers Limited 2016


Fig. 115. Tunable high-efficiency spin-filter. (a) Geometry of the spin-filter and the profile of the potential barrier U(xi). The two colored atoms on the lattice
emphasize the buckled structure. (b) Spin-polarization of the filter as a function of l0 in the constriction. The blue (green) line corresponds to the case of
applying potential barrier with a rectangular (smooth) shape. (c and d) are typical dispersion relations for the wide and the constriction regions,
respectively. Reproduced with permission from Ref. [128].
Ó Macmillan Publishers Limited 2016

Fig. 116. Y-shaped silicene spin-separator. A schematic Y-shape separator of silicene with three local gates (leads). The current separates into the gate 2 and
gate 3, respectively, carrying opposite spin/valley degrees of freedom. Reproduced with permission from Ref. [128].
J. Zhao et al. / Progress in Materials Science 83 (2016) 24–151 131

Ó American Physical Society 2016


Fig. 117. (a and b) Electronic band structures of freestanding and unsupported distorted silicene, respectively. (c and d) Electronic figure of merit at 300 K
as a function of chemical potential for freestanding and unsupported distorted silicene, respectively. Reproduced with permission from Ref.
[524].

Ó Royal Society of Chemistry 2016

Fig. 118. (a and b) The electrical conductance as a function chemical potential at 300 K for ASiNR and ZSiNR, respectively, with various ribbon widths. (c
and d) The Seebeck coefficient as a function chemical potential at 300 K for ASiNR and ZSiNR, respectively, with various ribbon widths. Reproduced with
permission from Ref. [526].

As a result, the supported silicene has an enhanced electronic figure of merit of 0.81 eV at room temperature [524]. However,
the lattice thermal conductivity of freestanding and supported silicene were reported to be 3–60 W/mK [75,79,80], at least
one order of magnitude larger than the electronic contribution. Therefore, the total thermoelectric figure of merit of silicene
is much smaller than the upper bound given by ZTe.
Nanostructuring silicene is an effective way to enhance the Seebeck coefficient and simultaneously reduce the lattice
thermal conductivity [525–527]. As discussed in the Session 4.5, silicene nanoribbons with the edges passivated by hydrogen
are all semiconductors [356]. For the armchair ribbons in the ground state, the band gap varies in an oscillating fashion with
the ribbon width w (Fig. 75a). In particular, ASiNRs with w = 7, 9, 10 have large band gaps. Accordingly, their electronic trans-
mission functions show clear stepwise features, and their Seebeck coefficients attain as much as 860 lV/K at room temper-
ature (Fig. 118a and c) [524,526]. For the ground-state zigzag ribbons, the band gap decreases monotonically as w increases,
and is generally smaller than that of ASiNRs (Fig. 75c). The Seebeck coefficient of ZSiNRs achieves a maximum of 200 lV/K
132 J. Zhao et al. / Progress in Materials Science 83 (2016) 24–151

Ó American Physical Society 2016


Fig. 119. The phonon thermal conductance as a function of temperature for (a) ASiNR and (b) ZSiNR, respectively. The thermoelectric figure of (c) ASiNR and
(d) ZSiNR as a function of ribbon width, respectively. Reproduced with permission from Ref. [524].

at room temperature (Fig. 118b and d). Due to the band gap opening, the power factor of silicene ribbons is greatly enhanced
compared to that of infinite silicene sheets [525–527].
For both ASiNRs and ZSiNRs, the phonon thermal conductance increases with temperature (Fig. 119a and b) [524]. The
ribbon with a smaller width has a lower phonon thermal conductance due to stronger phonon-boundary scattering. At room
temperature, the phonon contribution to the thermal conductance is of the same magnitude or slightly larger than the
electron contribution. Combining the electron and phonon thermal conductances, the total figure of merit for ASiNRs shows
an oscillatory variation with ribbon width. At room temperature, it reaches maximum values of 1.04 and 0.8 for w = 3 and 7,
respectively. For ZSiNRs, ZT decreases with increasing width, and it achieves the largest value of 0.6 with w = 3 for hole
doping (Fig. 119c and d).
As discussed previously in Section 4.5, ZSiNRs exhibit edge magnetism and antiferromagnetic state (Fig. 120a). By apply-
ing an external magnetic field, ZSiNRs can be easily switched from AFM to FM configuration (Fig. 120b), since the energy
difference between these two states is very small (only 0.02 eV for w = 5) [525]. The two magnetic states result in distinctly
different electronic band structures and transmission functions for ZSiNRs. The AFM configuration exhibits a semiconducting
Ó American Physical Society 2016

Fig. 120. Spin density of ZSiNR in the (a) antiferromagnetic, (b) ferromagnetic, and (c) antiparallel state. Reproduced with permission from Ref.
[525].
J. Zhao et al. / Progress in Materials Science 83 (2016) 24–151 133

Ó American Physical Society 2016


Fig. 121. (a) The electrical conductance, and (b) seebeck coefficient as a function of chemical potential for ZSiNR in the FM state, with a ribbon width of
w = 7 and at various temperatures. (c) Magnetoresistance and (d) magnetothermopower as a function of chemical potential for ZSiNR in a transition from
the AFM to FM state, with a ribbon width of w = 5 and at T = 90 K. Reproduced with permission from Ref. [525].

feature, and the band gap depends on the width of nanoribbons [524,526]. The FM configuration, on the other hand, behaves
like a metal with a constant transmission close to l = 0, and the transmission only weakly depends on the ribbon width
(Fig. 121a) [525]. As a consequence, the Seebeck coefficient of the FM state displays several peaks, the intensities of which
are remarkably smaller than that of the AFM configuration (Fig. 121b).
Due to the distinct electronic and transport properties of the two magnetic states, one can achieve a large magnetore-
sistance and magnetothermopower (MTP) in ZSiNRs by applying an external magnetic field [525]. MR and MTP are
defined as:
MR ¼ ðGFM  GAFM Þ=ðGFM þ GAFM Þ; ð31Þ
and
MTP ¼ ðSFM  SAFM Þ=ðjSFM j þ jSAFM jÞ; ð32Þ
respectively, where GFM (GAFM) is the electrical conductance of the FM (AFM) state, and SFM (SAFM) is the Seebeck coefficient of
the FM (AFM) state. At T = 90 K and for small doping levels, MR reaches 1 since GAFM is close to 0 due to existence of band gap
(Fig. 121c). For higher doping levels, negative magnetoresistance can be observed. MTP attains ±1 in wide regions near l = 0
(Fig. 121d), where SFM is negligibly small compared to SAFM. Therefore, by varying the magnetic configuration of ZSiNRs, one
can modulate not only the electrical conductance, but also the voltage generated by the temperature gradient.
ZSiNRs can also attain the AP configuration under a specific magnetic field. The AP state corresponds to a situation where
the two electrodes have opposite spin polarizations (Fig. 120c). The transmission of the AP configuration strongly depends on
the width of nanoribbons. For odd values of w (w = 5, 7), there is no band gap, and the transmission close to l = 0 resembles
that of the FM state. For w = 6, a band gap of about 0.4 eV appears; as a result, changing the magnetic configuration of ZSiNRs
from the AP to the FM state can lead to large magnetoresistance and magnetothermopower in the vicinity of l = 0 [525].
A temperature gradient can lead to not only charge accumulation but also spin accumulation at the ends of an open
system, when the spin channels are not mixed by spin-flip transitions [525,531]. The spin-dependent Seebeck coefficient
Sr (r = ", ;) corresponds to the voltage of spin-up and spin-down electrons induced by a temperature gradient DT as:
Sr ¼ DV r =DT ¼ L1r =jejTL0r ; ð33Þ
where Lnr is the Lorentz integral and involves the spin-dependent transmission function [531]. The charge (conventional)
thermopower Sc and spin thermopower Ss are defined as:
Sc ¼ ðS" þ S# Þ=2; ð34Þ
134 J. Zhao et al. / Progress in Materials Science 83 (2016) 24–151

Ó Royal Society of Chemistry 2016


Fig. 122. Spin-resolved (a) electronic band structure and (b) transmission function; (c) charge and spin Seebeck coefficient, (d) electron thermal
conductance, and (e) charge and spin figure of merit at T = 90 K, for ZSiNR with the 1H–0H type of edges. The inset in (d) shows the phonon thermal
conductance as a function of temperature. The inset in (e) shows the maxima of charge and spin figure of merit at negative chemical potentials as a function
of temperature. Reproduced with permission from Ref. [532].

and

Ss ¼ ðS"  S# Þ=2; ð35Þ

respectively. For materials with dimension shorter than the spin-flip length, and having spin-dependent transport proper-
ties, spin Seebeck effect can be observed [525,531]. The conventional thermopower can be affected as well due to spin
separation and accumulation. Under such circumstances, one can introduce the charge (conventional) ZTc and spin thermo-
electric figure of merit ZTs as:

ZTc ¼ S2c GT=j; ð36Þ

and

ZTs ¼ S2s jGs jT=j; ð37Þ

respectively, where G = G" + G; is the total electrical conductance, Gs = G"  G; is the spin conductance, and j is the total
thermal conductance from electrons and phonons [531].
ZSiNRs may have non-degenerate electron bands from the two spin channels, depending on the magnetic configuration,
hydrogenation level of the edges, impurities and defects [531–535]. Prominent spin thermoelectric effects have been
observed in various ZSiNRs [525,531–533,536]. One representative is the ZSiNRs with asymmetrically hydrogenated edges,
where one edge is mono-hydrogenated and another side is either bare or di-hydrogenated (referred to as 1H–0H and 1H–2H,
respectively). The ZSiNRs with 1H–0H and 1H–2H adopt the ferromagnetic configuration as ground state. Both of them are
semiconductors, and show very peculiar band structures, i.e., electronic states from the spin-up channel exist in the spin-
down gap near the Fermi level and vice versa (Fig. 122a and b). In other words, there are regions of chemical potential, where
one of the spin channels is blocked for transport while the other is conductive. This leads to large spin conductance with
±100% polarization close to the Fermi energy. Moreover, spin thermopower attains as large as 1.4 mK/V due to the sharp
change of transmission near the gap edge (Fig. 122c). As a result, these systems exhibit remarkably enhanced conventional
and spin thermoelectric figure of merit, e.g., up to 9 at T = 90 K (Fig. 122d).
Apart from edge hydrogenation, introducing impurities and vacancies can break the transversal symmetry of ZSiNRs and
remove the degeneracy of spin channels, giving rise to very interesting spin Seebeck effects [531,533–535]. A ZSiNRs hetero-
junction with source and drain made up of mono- and di-hydrogenated ZSiNRs, respectively, was proposed to generate pure
thermal spin current and induce various spin caloritronic phenomena by modulating the temperature [537]. A recent study
showed that the topological properties of ZSiNRs have significant influence on the thermoelectric phenomena of the system.
J. Zhao et al. / Progress in Materials Science 83 (2016) 24–151 135

Both a staggered exchange field and a normal electric filed can separately open a gap in the edge states of ZSiNRs; and the
gap is spin-dependent by the interplay of the two fields, leading to nonzero conventional and spin thermopower close to the
gap edge [538].
As demonstrated above, silicene-based materials possess spectacular thermoelectric and spin thermoelectric properties,
which can be tuned by controlling the dimension and structure of the systems, functionalization of edges, doping, as well as
by applying external electric and magnetic fields. Hence, they are very promising candidates for nanoelectronic, spintronic,
and thermoelectric devices.

6.2. Chemical sensor

The peculiar electronic properties of silicene render its high sensitivity to certain gaseous molecule and chemical species;
thus silicene based chemical sensor can be developed. Feng et al. [539] systematically investigated the adsorption of several
common gas molecules (CO, NO, NO2, O2, CO2, NH3, and SO2) on silicene sheet by DFT calculations. They found that silicene is
a promising candidate as the sensor of NO and NH3 since these two molecules can be chemically adsorbed on silicene with
moderate adsorption energies (0.35 eV for NO and 0.6 eV for NH3), certain amount of charge transfer (0.08 e for NO and
0.3 e for NH3), and opening of a small band gap (0.05–0.08 eV). Meanwhile, silicene may serve as a metal-free catalyst for
activating and catalyzing NO2, O2, and SO2 gases owing to the large adsorption energies (>1 eV). Furthermore, the Stone–
Wales defect and Ag(1 1 1) substrate can enhance the chemical reactivity of silicene in terms of binding energy and charge
transfer. Recently, the adsorption characteristics of NO molecule on pristine, Al- and P-substituted silicene nanosheets of
finite size (Si30H14) have been theoretically discussed in terms of adsorption energy, HOMO–LUMO gap and Mulliken charge
transfer [540].
An independent first-principles study by Yang’s group [541] also demonstrate the potential application of silicene as a
highly sensitive molecule sensor for NH3, NO, and NO2 molecules. All these three gas molecules chemically adsorb on silicene
with distinct charge transfer from silicene to molecules, opening a tunable band gap at the Dirac point of silicene. Most inter-
estingly, the calculated charge carrier concentrations of NO2-adsorbed silicene are three orders of magnitude larger than the
intrinsic charge carrier concentration of pristine graphene at room temperature, suggesting an effective enhancement of the
hole conductivity in silicene that is beneficial for nanoelectronic devices. Overall speaking, the strong binding of gas mole-
cules to silicene and the versatile electronic properties of the adsorbed silicene sheets may lead to potential applications in
gas detection and catalysis.
Using the non-equilibrium Green’s function method, Osborn and Farajian [542] demonstrated that the semiconducting
silicene nanoribbons may serve as carbon monoxide nanosensors with molecular resolution. With weak chemisorption of
a single CO molecule, the quantum conductance of silicene nanoribbon is detectably modified due to the charge transfer
from CO to the silicene (about 0.4 |e|) and the adsorption-induced structural deformation. The effects of atmospheric gases
including nitrogen, oxygen, carbon dioxide, and water were also considered. Adsorption of each individual environmental
gas molecule modulates the electron conduction of pristine silicene nanoribbons in a differentiable manner from CO.
In addition to the gas sensors discussed above, silicene-based DNA nucleobase sensors have been theoretically proposed.
Using DFT combined with quantum scattering theory, Sadeghi et al. [544] computed the electrical transport properties of
silicene nanoribbon with a nanopore of 1.7 nm diameter in the absence and presence of nucleobases. It was shown that
the electrical conductance of silicene nanoribbons with a nanopore is selectively sensitive to the translocation of DNA nucle-
obases through the pore. Later, Amorim and Scheicher [543] performed a comprehensive DFT calculation on the interaction
between DNA nucleobases and silicene sheet (Fig. 123) and found that adenine and thymine are physisorbed on silicene,
whereas cytosine and guanine are weakly chemisorbed through the formation of a Si–O bond. Most importantly, the electron
Ó IOP Publishing 2016

Fig. 123. (Left) Top views and side views of the fully relaxed geometries for the four different nucleobases (A, C, G, and T) physisorbed on a buckled silicene
sheet. The hexagonal network of Si atoms is drawn in salmon color, C atoms are represented as green spheres, N atoms in blue, O atoms in red, and H atoms
in white. (Right) The calculated sensitivity of the hypothetical silicene sequencing device with respect to the four different nucleobases (A, C, G, and T) for
two different gate voltages (1.26 V and +1.00 V). Reproduced with permission from Ref. [543].
136 J. Zhao et al. / Progress in Materials Science 83 (2016) 24–151

transmission of each nucleobase differs from that of pristine silicene in a characteristic manner. The sensitivity histograms in
Fig. 123 demonstrate that a distinction of all four nucleobases is possible by combining the conductance data from measure-
ments at two gate voltages (1.26 V and +1.00 V). Definitely, these pioneer studies open the avenues towards fast, cheap,
and portable DNA sequencing with silicene-based electrical sensors, which have immediate advantage of being compatible
with existing silicon CMOS technologies.

6.3. Hydrogen storage

Similar to its carbon analogue (i.e., graphene), silicene has been considered as a promising hydrogen storage material. In
general, hydrogen can be stored in the form of chemical hydrides (namely, chemical storage) or physisorbed gaseous H2
molecules (namely, physical storage). Using first-principles calculations, Jose and Datta [545] have explored the chemical
hydrogen storage properties of a number of silicene nanoflakes. After fully hydrogenation on both upper/lower sides and
at the edge dangling Si atoms, the weight percent of hydrogen ranges from 6.6% (Si6H12) to 4.5% (Si70H92). Recent experimen-
tal by Wu’s group [39] demonstrated that the fully hydrogenated silicene sheet on Ag(1 1 1) substrate can be completely
restored to the pristine silicene by annealing to a moderate temperature of about 450 K. Such easily reversible hydrogenation
of monolayer silicene suggests that silicene may be useful for controllable chemical storage of hydrogen.
By itself, silicene interacts weakly with H2 molecule and cannot be used for physical storage of hydrogen gas. Alterna-
tively, there have been considerable computational efforts on designing physical hydrogen storage material using metal dec-
orated silicene or silicane. All those calculations considered a (2  2) supercell with eight Si atoms, whereas silicon atoms
have been hydrogenated to form silicane in some cases. As a prototype model system, the structures of Na-decorated silicene
(NaSi8) adsorbed with different amount of hydrogen are shown in Fig. 124 [546]. Since each Na atom can store seven H2
molecules, the maximum gravimetric density of H2 reaches 9.40 wt.% when Na atoms sit on two opposite sides of silicene.
The current theoretical results on physical storage of hydrogen with metal-decorated silicene/silicane are summarized in
Table 9. To achieve hydrogen uptake/release at moderate conditions, the desired average binding energy of each H2 molecule

Fig. 124. Schematics of adsorption of H2 molecules on single Na atom decorated (2  2) unit cell of silicene (NaSi8). Panels (a–f) correspond to the number
of H2 varying from 2 to 7. For each panels, top view and side view are displayed. Reproduced with permission from Ref. [546]. Copyright 2016, Elsevier B.V.

Table 9
Maximum number of H2 per adsorbed on each metal atom (Nmax), maximum gravimetric density of H2 (v), average adsorption energy (Ead) per H2 at saturation
concentration.

Stoichiometry Nmax v (wt.%) Ead (eV) Reference


Li2Si8 5 7.75 0.46 [550]
Na2Si8 5 6.90 0.48 [550]
Li2Si8H6 4 6.30 0.37 [548]
Na2Si8H6 4 5.40 0.39 [548]
Li2Si8 4 6.35 0.26 [551]
K2Si8H8 5 6.13 0.133 [552]
Mg2Si8 6 8.10 – [549]
Mg2Si8H6 6 7.95 – [549]
CaSi8 9 6.40 0.19 [553]
Na2Si8 7 9.40 – [546]
K2Si8 6 7.31 0.18 [546]
Ca2Si8 7 – 0.24 [546]
J. Zhao et al. / Progress in Materials Science 83 (2016) 24–151 137

Ó American Chemical Society 2016


Fig. 125. Atomic structures corresponding to lithiated single-layer (SL) and double-layer (DL) silicene, LixSi1x of varying Li content, x. The unit cell is
outlined in each top view; top and side views are shown separately. Reproduced with permission from Ref. [554].

should be in the range of 0.2–0.4 eV [547], while the DOE target of gravimetric density for on-broad applications is 5.5 wt.%
by 2017 [548]. As seen in Table 9, most of the predicted maximum gravimetric densities of silicene/silicane appropriately
decorated with light metals (Li, Na, K, Mg, Ca) meet the DOE target, and the H2 binding energies fall in the desired range.
For a given metal atom, the specific strength of metal–H2 interaction and saturation number of adsorbed hydrogen molecule
depend on the theoretical method used and vary in different literature.
In addition, the hydrogen storage properties of metal-decorated silicene can be further tuned by external strain or electric
field. For instance, Ahuja’s group found a drastic increase (by about 80%) in the binding energy of Mg adatoms on the silicene/
silicane monolayer sheet under the biaxial symmetric strain of 10%, which ensures the uniform distribution of Mg atoms
over the substrates [549]. Under a perpendicular electric field (F), the average binding energy per H2 (0.19 eV) on Ca atom
supported on monolayer or bilayer silicene can be enhanced to 0.37 eV (F = 0.004 a.u.) or reduced to 0.02 eV (F = 0.004 a.u.).
As a consequence, external electric field can be utilized as a switch to control the adsorption and desorption of the hydrogen
molecules on Ca-decorated silicene systems.

6.4. Electrode material for Li battery

To meet the growing demand of portable electronic products and electric vehicles, Li-ion batteries (LIBs) have been devel-
oped for future energy storage and utilization. However, the energy density of LIBs is restricted by the capacity limits of the
electrode materials, e.g., 372 mA h/g for graphite anode. Owing to the very high theoretical specific capacity (4200 mA h/g),
Si-based anodes have been considered as an alternative to the conventional graphite anode. Especially, as monolayer of
silicon atoms, silicene with large surface area could serve as high-capacity host of Li in LIB.
Using DFT calculations, Tritsaris et al. [554] investigated the interaction of Li with Si in model electrodes of freestanding
single-layer (SL) and double-layer (DL) silicene, which are displayed in Fig. 125. According to their theoretical results, silicene
is suitable for Li-ion storage due to the following advantages: (i) much higher theoretical specific charge capacity than
graphite (954 and 715 mA h/g for the SL and DL silicene respectively); (ii) large binding energy of Li (2.2 eV per Li atom)
that is nearly irrelevant to Li content and silicene thickness; (iii) relatively low barriers for Li diffusion (typically less than
0.6 eV); (iv) small change in the effective volume (13% for SL and 24% for DL structures) during lithiation and delithiation
cycles.
By means of DFT computations, Deng et al. [555] investigated lithium adsorption and diffusion on AC and ZZ silicene
nanoribbons of about 4 nm wide. It was found that monolayer silicene edges provide the strongest binding energy with
Li among known structural forms of silicon (2.42 eV for ZZ and 2.20 eV for AC, respectively). The energy barriers for Li
diffusion on monolayer silicene nanoribbons are very low: 0.14 eV for ZZ and 0.26 eV for AC. Therefore, silicene nanoribbons
as anode materials exhibit good reactivity with charge carrying Li ions and allow for their fast transport; both are beneficial
for developing LIBs with large energy densities and high rate capabilities.
In addition to pristine silicene, Chen’s group [556] proposed that the boron-substituted silicene, i.e., BSi3 with planar
geometry is a superior candidate material for anodes of LIBs. Compared to pristine silicene, BSi3 has even higher theoretical
charge capacity (1410 and 846 mA h/g for single- and double-layer, respectively) and better electronic conductivity due to its
intrinsically metallic nature. Moreover, the relatively low average open-circuit voltage (0.43 V for SL and 1.38 V for DL) and
low diffusion barriers of Li (typically less than 0.4 and 0.6 eV for SL and DL, respectively) suggest that BSi3 silicene could serve
as a promising high capacity and fast charge/discharge rate anode material for LIBs.
138 J. Zhao et al. / Progress in Materials Science 83 (2016) 24–151

Fig. 126. Calculated free energy diagram for the ORR/OER on silicene at zero potential (U = 0 V), at the maximum discharge potential (UDC = 1.36 V), at the
minimum charge potential (UC = 6.84 V), and at the theoretical equilibrium potential (U0 = 2.76 V). Reproduced with permission from Ref. [558]. Copyright
2016, Elsevier B.V.

Considering the limited resource of lithium and the abundant deposit of sodium in earth crust, Na-ion batteries (NIBs)
have attracted wide attentions as alternative of LIBs for large-scale energy storage in smart electric grids. Kulish et al.
[557] performed first-principles calculations on the insertion and diffusion of Na and Li in layered silicon materials
(i.e., polysilane and silicene) as potential anode materials for Na-ion batteries. They found that insertion of Na atom in
the layered polysilane and silicene is exothermic with negative (favorable) binding energies of 0.57 and 0.32 eV, respec-
tively, compared to that of +0.6 eV in bulk silicon (unfavorable). Furthermore, the energy barrier for Na diffusion is reduced
from 1.06 eV in bulk silicon to 0.41 eV in layered polysilane and 0.14 eV in silicene, corresponding to much faster diffusion
rates. Therefore, the major shortcomings of the bulk Si solid, i.e., unfavorable Na insertion energetics and slow Na kinetics,
can be avoided, suggesting that these layered silicon structures are promising anode materials for NIBs.
Compared to the conventional Li-ion batteries, Li air batteries depict super promising prospects due to its extremely high
theoretical energy density. Based on DFT calculations, Chung’s group [558] demonstrated the potential usage of silicene as a
carbon-free cathode structure for Li-air batteries. Fig. 126 shows the calculated free energy diagram for the oxygen reduction
reactions (ORR) and oxygen evolution reactions (OER) on silicene, illustrating the four reactions steps with the optimized
intermediates: LiO⁄2, Li2O⁄2, and (Li2O)2. It can be observed that an ORR/OER occurs on the pristine form of silicene without
any additional catalysts, in contrast to graphene which shows catalytic activity for ORR/OER only through the defect sites.
The higher adsorption energies for the ORR intermediates on silicene are closely related to not only the charge transfer
and ionic bonding between the silicene and the ORR intermediates, but also the structural changes of the silicene structures
into a more sp3-like form of structure. In particular, silicene as a cathode material for Li air batteries might be able to enhance
the reversible cycle life without producing Li2CO3-like species.

7. Summary and prospects

As we mentioned above, silicene presents a kind of unique 2D material that possesses many outstanding properties and
holds promise for a variety of applications. Our current knowledge on the fundamental properties and the possible applica-
tions of silicene can be briefly summarized into the following points.

(1) The electronic properties of silicene resemble those of graphene, e.g., Dirac cone, high Fermi velocity, and high carrier
mobility. However, due to the related low-buckled geometry with partial sp3 hybridization, the intrinsic properties of
silicene differ from graphene with pure sp2 hybridization in some aspects, e.g., a larger SOC gap of 1.55 meV, and a
tunable band gap, an easier valley polarization, and much smaller thermal conductivity of 3–65 W/mK. In particular,
the stronger SOC effect in silicene with regard to graphene would result in quantum spin Hall effect that may be
realized in experimentally accessible temperature. Besides QSHE, silicene may also host other exotic quantum states,
such as quantum valley Hall effect, quantum anomalous Hall effect, phonon-mediated superconductor, and topolog-
ical superconductor.
(2) Experimental efforts on silicene have generated significant results in the last few years. Monolayer silicene has been
successfully prepared on a variety of substrates: Ag(1 1 1), ZrB2(0 0 0 1), and other substrates. The formation mecha-
nism of silicene on different substrates, especially the various superstructures that are induced by the unique buckling
of the silicene lattice, has been much understood based on experimental observations and theoretical calculations.
Some silicene derivatives, such as silicene nanoribbon, multilayer silicene, hydrogenated silicene, have been also
reported, which renders more possibility in the application of silicene. Although there is still a long journey to the real-
ization of a usable silicene device, significant progresses have been made toward this direction, for examples, methods
J. Zhao et al. / Progress in Materials Science 83 (2016) 24–151 139

to protect the silicene surface from oxidation, such as Al2O3 protection layers, were testified, which results in the
demonstration of the first silicene-based FET device. In the future, the prospect of silicene application still mainly
relies on the overcoming of two difficulties: the stability of silicene at ambient condition and the finding of a suitable
insulating substrate.
(3) The physical and chemical properties of pristine silicene can be further modulated by various ways of modification.
Hydrogenation, halogenation and oxidization are all able to open band gap and transform silicene from semimetal
to semiconductor or insulator, whereas half hydrogenation or half-fluorination of silicene would further induce inter-
esting magnetic behaviors. Adsorption or intercalation of metal atoms is also an effective approach to open band gap
and introduce abundant magnetic and topological states in silicene. Owing to the quantum confinement effect,
silicene nanoribbons and nanoflakes exhibit novel physical properties different from the infinite sheet, e.g., width-
oscillating band gap, spin-polarized half metallicity. Moreover, defect and strain in realistic silicene materials are
not only inevitable, but also provide new opportunities in engineering the electronic and magnetic properties of
silicene. Substitutional doped silicene and hereto silicene sheets can be metallic, semimetallic or semiconducting,
largely extending the diversity of silicene materials. The interaction of silicene with substrate or other 2D layer mate-
rial would affect the electronic properties of silicene and lead to novel heterostructures with tunable band gap under
electric field.
(4) Opening a sizable band gap without degrading its electronic properties is probably one of the most urgent issues for
silicene application. This goal can be achieved more easily in silicene compared with graphene due to a tunable band
gap. Applying a vertical electric field, digging holes in silicene to create a nanomesh, and single-side adsorption of
atoms or molecules are able to open a band gap without destructing the electronic properties of silicene. Experimen-
tally, the reported Ion/Ioff is 10 with a band gap of 0.21 eV in the fabricated silicene FET. By taking full advantage of the
theoretically proposed band gap opening strategies, the performance of silicene FET could be greatly improved. As for
the silicene spintronics, half-metallicity can be achieved in zigzag silicene nanoribbons with a transverse electric filed/
asymmetric edge modifications or semihydrogenated silicene with an out-of-plane gate voltage. Quantum transport
simulations of spin-filter and spin FET based on these half-metallic silicene nanoribbons show high performances.
Giant magnetoresistance can be obtained in zigzag silicene nanoribbons either by utilizing the switch between differ-
ent magnetic states or the symmetry-dependent transport property. In addition, silicene and its derivatives serve as
candidate materials for many other potential applications, e.g., thermoelectric device, chemical sensor, hydrogen stor-
age, electrode material for Li batteries.

After the experimental synthesis of large-scale silicene sheet in 2012, there have been many great achievements in this
field, especially the recent fabrication of silicene FET. Even so, the research of silicene is still at its early stage remaining with
great challenges: the preparation of silicene usually required ultrahigh vacuum environment, suitable substrates are difficult
to find, and the silicene surface is sensitive to oxidation. Certainly, making silicene devices are far more difficult than the case
of graphene. The existing difficulties and debates of silicene research, however, should not stop the community from moving
forward. Just think about the important fact: the silicon industry has been developed for decades and dominated our life, but
a hexagonal silicon sheet, with graphene structure, has just been found in the last few years. One can imagine how strong
impact it would have once the intriguing properties of silicene can be integrated to the existing silicon-based devices.
To gain deeper insights into the physical and chemical properties of silicene and to eventually utilize silicene in the future
nanoelectronic circuits and other 2D devices, we believe that the following critical issues have to be addressed in silicene
research from both experimental and theoretical points of view.

(1) In the experimental synthesis of silicene, in particular on Ag(1 1 1) substrate, there are co-existence of several super-
structures. Moreover, the maximum domain size of monolayer silicene is only less than 1 lm. Therefore, deeper
understanding on the growth mechanism of silicene at the atomic level is desirable to achieve larger-scale high-
quality and uniform silicene sheets, desirably up to centimeter scale like graphene.
(2) Owing to the complicated superstructures and the resulting intricate STM patterns of various silicene sheets, it is nec-
essary but quite challenging to identify the structural defects (e.g., Stone–Wales rotation, single vacancy, double
vacancy) in the silicene samples grown on substrates. Moreover, it is important to reduce the defect density during
fabrication process in order to avoid the defect-induced degradation of carrier mobility of silicene.
(3) Most of the fascinating properties of silicene are associated with existence of the Dirac cone. Keeping the Dirac cone is
critical for observation of the predicted fascinating properties. So far, there have been many controversies about
whether Dirac cones exist for monolayer silicene sheets on Ag substrates from experimental side though most of
the theoretical works suggest that the Dirac cone of silicene is absent on Ag substrates due to the strong band
hybridization. Further spectroscopic studies of silicene on Ag substrates are still required to finally resolve this issue.
(4) For device application, it is critically important to transfer as-grown monolayer silicene sheets from metallic substrate
to insulating one, as the treatment of graphene. However, the much enhanced adhesion between silicene and metallic
substrates renders this transfer scheme rather difficulty. To directly grow silicene on semiconducting or insulating
materials is a good alternative, especially in light of the success growth of silicene on semiconducting MoS2. According
to the theoretical works, growth of silicene on a semiconducting or insulating substrate with a small lattice mismatch
should be an endeavor direction because the Dirac cone can be well kept.
140 J. Zhao et al. / Progress in Materials Science 83 (2016) 24–151

(5) Compared to the abundant spectroscopic data of graphene and other 2D materials like transition-metal dichalco-
genides, currently the spectroscopic characterizations of silicene samples are still rather limited and deserve more
attentions.
(6) What is the difference between multilayer silicene and ultrathin layer of bulk silicon of diamond phase? How to
distinguish them experimentally? In contrast to the rapid progress on experiments, there is barely any theoretical
simulation on multilayer silicene on substrates, although the freestanding multilayer silicene sheets have been
comprehensively explored by ab initio calculations.
(7) According to theoretical predictions, there are many possible ways to open a band gap in silicene and to modulate the
gap amplitude. Magnetism and half metallicity are also predicted in silicene nanoribbons. How to realize these theo-
retical predictions in experiments, however, remains a big challenge. In addition, theoretical calculations have also
revealed many novel quantum phenomena in silicene, such as QSHE, QAHE, and topological superconductivity. But
it also is a great challenge to observe these novel physical properties in experiments.
(8) Although there have already been some pioneering efforts, innovation designs of silicene based devices utilizing its
fantastic quantum properties are still high desirable to guide and motivate further experiments.
(9) In recent experiments, there have been some progress in hydrogenation and metal doping (or coating) of silicene.
More experiments on chemical modification (e.g., fluorination, alkali metal intercalation) of silicene are still antici-
pated to tailor the physical and chemical properties of silicene and to verify the earlier theoretical predictions.
(10) The high reactivity of silicene not only leads to strong adhesion with substrate but also results in a rapid degradation
of silicene device when exposed to ambient air. Therefore, feasible silicene device must seek a shield from ambient air.
(11) It would be exciting to fabricate silicene based heterostructures via either in-plane integration or inter-layer stacking
with other 2D materials like graphene, h-BN, germanene, transition metal dichalcogenide nanosheets. Note that the
measured carrier mobility in the present silicene FET is one order of magnitude smaller than that expected from
the theoretical calculation, probably due to the roughness of substrate. Experimentally, sandwiching few layer phos-
phorene by planar hexagonal BN effectively prevents the degradation of phosphorene and simultaneously improves
the carrier mobility significantly [559]. Sandwiching silicene by hexagonal BN layers is expected to prevent the rapid
degradation of silicene and improve the carrier mobility as effectively as sandwiching phosphorene.

Finally, it is noteworthy that silicene is not the only member of 2D monolayer materials featured with Dirac cone. In the
periodic table, monolayer sheets of other group IV elements, such as germanene and stanene, are also attractive because of
their stronger effect of spin–orbit coupling. In particular, stanene and its derivatives were theoretically predicted to be 2D
topological insulators with a large band gap [560] and exhibit the near-room-temperature quantum anomalous Hall effect
[561]. Recent experiments have successfully fabricated monolayer or few-layer germanene on Au(1 1 1) [562,563], Pt(1 1 1)
[564] and Al(1 1 1) surfaces [256,257], as well as epitaxial hexagonal AlN buffer layer on Ag(1 1 1) substrate [565]. Epitaxial
growth of 2D stanene film on Bi2Te3(1 1 1) substrates by MBE was also reported [566]. Besides the commonly used epitaxial
grow approaches, multilayered germanane sheets of millimeter-scale have been obtained from the topochemical deinterca-
lation of CaGe2, and these sheets can be further mechanically exfoliated as single and few layers onto SiO2/Si surfaces [567].
In addition to the counterparts of silicene made up of group IV elements, monolayer materials with the Dirac cone are also
predicted in a novel 2D boron structure with nonzero thickness [568] and single-layer Hf film with honeycomb structure
(namely, hafnene) [569]. The latter one has also been experimentally synthesized on the Ir(1 1 1) substrate. All these exciting
progress demonstrate many opportunities in the design and synthesis of 2D materials with peculiar electronic band struc-
tures featured with the Dirac cone. Therefore, the successive research of silicene and the related nanostructures would not
only move forward silicene-based devices and materials but also lead to discovery of other novel 2D materials. In this regard,
we believe that silicene and its cousins like germanene and stanene have a bright future.

Acknowledgments

This work was supported by the National Natural Science Foundation of China (No. 11574040, 11334011, 11504041,
11574029, 11274016, 11474012, 11134005, 11225418), the National Basic Research Program of China
(No. 2012CB619304, 2013CB932604, 2014CB920903), the Fundamental Research Funds for the Central Universities of China
(No. DUT16LAB01), and the Specialized Research Fund for the Doctoral Program of Higher Education of China (Grants
No. 20121101110046).

References

[1] Novoselov KS, Geim AK, Morozov SV, Jiang D, Zhang Y, Dubonos SV, et al. Electric field effect in atomically thin carbon films. Science 2004;306:666–9.
[2] Xu M, Liang T, Shi M, Chen H. Graphene-like two-dimensional materials. Chem Rev 2013;113:3766–98.
[3] Tang Q, Zhou Z. Graphene-analogous low-dimensional materials. Prog Mater Sci 2013;58:1244–315.
[4] Miro P, Audiffred M, Heine T. An atlas of two-dimensional materials. Chem Soc Rev 2014;43:6537–54.
[5] Balendhran S, Walia S, Nili H, Sriram S, Bhaskaran M. Elemental analogues of graphene: silicene, germanene, stanene, and phosphorene. Small
2015;11:640–52.
[6] Takeda K, Shiraishi K. Theoretical possibility of stage corrugation in Si and Ge analogs of graphite. Phys Rev B 1994;50:14916–22.
[7] Guzmán-Verri GG, Lew Yan Voon LC. Electronic structure of silicon-based nanostructures. Phys Rev B 2007;76:075131.
J. Zhao et al. / Progress in Materials Science 83 (2016) 24–151 141

[8] Cahangirov S, Topsakal M, Aktürk E, Sßahin H, Ciraci S. Two- and one-dimensional honeycomb structures of silicon and germanium. Phys Rev Lett
2009;102:236804.
[9] Lebègue S, Eriksson O. Electronic structure of two-dimensional crystals from ab initio theory. Phys Rev B 2009;79:115409.
[10] Liu C-C, Feng W, Yao Y. Quantum spin Hall effect in silicene and two-dimensional germanium. Phys Rev Lett 2011;107:076802.
[11] Novoselov KS, Jiang D, Schedin F, Booth TJ, Khotkevich VV, Morozov SV, et al. Two-dimensional atomic crystals. Proc Natl Acad Sci USA
2005;102:10451–3.
[12] Zhao J, Liu L, Li F. Graphene oxide: physics and applications. Springer; 2014.
[13] Nakano H, Mitsuoka T, Harada M, Horibuchi K, Nozaki H, Takahashi N, et al. Soft synthesis of single-crystal silicon monolayer sheets. Angew Chem Int
Ed 2006;45:6303–6.
[14] Léandri C, Oughaddou H, Aufray B, Gay JM, Le Lay G, Ranguis A, et al. Growth of Si nanostructures on Ag(0 0 1). Surf Sci 2007;601:262–7.
[15] Aufray B, Kara A, Vizzini S, Oughaddou H, Léandri C, Ealet B, et al. Graphene-like silicon nanoribbons on Ag(1 1 0): a possible formation of silicene. Appl
Phys Lett 2010;96:183102.
[16] De Padova P, Quaresima C, Olivieri B, Perfetti P, Le Lay G. Sp2-like hybridization of silicon valence orbitals in silicene nanoribbons. Appl Phys Lett
2011;98:081909.
[17] De Padova P, Quaresima C, Ottaviani C, Sheverdyaeva PM, Moras P, Carbone C, et al. Evidence of graphene-like electronic signature in silicene
nanoribbons. Appl Phys Lett 2010;96:261905.
[18] Rachid Tchalala M, Enriquez H, Mayne AJ, Kara A, Roth S, Silly MG, et al. Formation of one-dimensional self-assembled silicon nanoribbons on Au
(1 1 0)-(2  1). Appl Phys Lett 2013;102:083107.
[19] Vogt P, De Padova P, Quaresima C, Avila J, Frantzeskakis E, Asensio MC, et al. Silicene: compelling experimental evidence for graphene like two-
dimensional silicon. Phys Rev Lett 2012;108:155501.
[20] Feng B, Ding Z, Meng S, Yao Y, He X, Cheng P, et al. Evidence of silicene in honeycomb structures of silicon on Ag(1 1 1). Nano Lett 2012;12:3507–11.
[21] Lin C-L, Arafune R, Kawahara K, Tsukahara N, Minamitani E, Kim Y, et al. Structure of silicene grown on Ag(1 1 1). Appl Phys Exp 2012;5:045802.
[22] Jamgotchian H, Colignon Y, Hamzaoui N, Ealet B, Hoarau JY, Aufray B, et al. Growth of silicene layers on Ag(1 1 1): unexpected effect of the substrate
temperature. J Phys: Condens Matter 2012;24:172001.
[23] Chiappe D, Grazianetti C, Tallarida G, Fanciulli M, Molle A. Local electronic properties of corrugated silicene phases. Adv Mater 2012;24:5088–93.
[24] Gao J, Zhao J. Initial geometries, interaction mechanism and high stability of silicene on Ag(1 1 1) surface. Sci Rep 2012;2:861.
[25] De Padova P, Vogt P, Resta A, Avila J, Razado-Colambo I, Quaresima C, et al. Evidence of Dirac fermions in multilayer silicene. Appl Phys Lett
2013;102:163106.
[26] De Padova P, Avila J, Resta A, Razado-Colambo I, Quaresima C, Ottaviani C, et al. The quasiparticle band dispersion in epitaxial multilayer silicene. J
Phys: Condens Matter 2013;25:382202.
[27] Fleurence AFR, Ozaki T, Kawai H, Wang Y, Yukiko YT. Experimental evidence for epitaxial silicene on diboride thin films. Phys Rev Lett
2012;108:245501.
[28] Aizawa T, Suehara S, Otani S. Silicene on zirconium carbide (1 1 1). J Phys Chem C 2014;118:23049–57.
[29] Meng L, Wang Y, Zhang L, Du S, Wu R, Li L, et al. Buckled silicene formation on Ir(1 1 1). Nano Lett 2013;13:685–90.
[30] Chiappe D, Scalise E, Cinquanta E, Grazianetti C, van den Broek B, Fanciulli M, et al. Two-dimensional Si nanosheets with local hexagonal structure on
a MoS2 surface. Adv Mater 2014;26:2096–101.
[31] Ni Z, Liu Q, Tang K, Zheng J, Zhou J, Qin R, et al. Tunable bandgap in silicene and germanene. Nano Lett 2012;12:113–8.
[32] Drummond ND, Zólyomi V, Fal’ko VI. Electrically tunable band gap in silicene. Phys Rev B 2012;85:075423.
[33] Ni Z, Zhong H, Jiang X, Quhe R, Luo G, Wang Y, et al. Tunable band gap and doping type in silicene by surface adsorption: towards tunneling
transistors. Nanoscale 2014;6:7609–18.
[34] Quhe R, Fei R, Liu Q, Zheng J, Li H, Xu C, et al. Tunable and sizable band gap in silicene by surface adsorption. Sci Rep 2012;2:853.
[35] Liu H, Han N, Zhao J. Band gap opening in bilayer silicene by alkali metal intercalation. J Phys: Condens Matter 2014;26:475303.
[36] Houssa M, Pourtois G, Afanas’ev VV, Stesmans A. Can silicon behave like graphene? A first-principles study. Appl Phys Lett 2010;97:112106.
[37] Lew Yan Voon LC, Sandberg E, Aga RS, Farajian AA. Hydrogen compounds of group-IV nanosheets. Appl Phys Lett 2010;97:163114.
[38] Wang R, Pi X, Ni Z, Liu Y, Lin S, Xu M, et al. Silicene oxides: formation, structures and electronic properties. Sci Rep 2013;3:3507.
[39] Qiu J, Fu H, Xu Y, Oreshkin AI, Shao T, Li H, et al. Ordered and reversible hydrogenation of silicene. Phys Rev Lett 2015;114:126101.
[40] Du Y, Zhuang J, Liu H, Xu X, Eilers S, Wu K, et al. Tuning the band gap in silicene by oxidation. ACS Nano 2014;8:10019–25.
[41] Ding Y, Ni J. Electronic structures of silicon nanoribbons. Appl Phys Lett 2009;95:083115.
[42] Tao L, Cinquanta E, Chiappe D, Grazianetti C, Fanciulli M, Dubey M, et al. Silicene field-effect transistors operating at room temperature. Nat
Nanotechnol 2015;10:227–31.
[43] Liu C-C, Jiang H, Yao Y. Low-energy effective Hamiltonian involving spin–orbit coupling in silicene and two-dimensional germanium and tin. Phys Rev
B 2011;84:195430.
[44] Chen L, Feng B, Wu K. Observation of a possible superconducting gap in silicene on Ag(1 1 1) surface. Appl Phys Lett 2013;102:081602.
[45] Kara A, Enriquez H, Seitsonen AP, Lew Yan Voon LC, Vizzini S, Aufray B, et al. A review on silicene—new candidate for electronics. Surf Sci Rep
2012;67:1–18.
[46] De Padova P, Perfetti P, Olivieri B, Quaresima C, Ottaviani C, Le Lay G. 1D graphene-like silicon systems: silicene nano-ribbons. J Phys: Condens Matter
2012;24:223001.
[47] Deepthi J, Ayan D. Structures and chemical properties of silicene: unlike graphene. Acc Chem Res 2013;47:593–602.
[48] Yan Voon LCL, Guzmán-Verri GG. Is silicene the next graphene? MRS Bull 2014;39:366–73.
[49] Yamada-Takamura Y, Friedlein R. Progress in the materials science of silicene. Sci Technol Adv Mater 2014;15:064404.
[50] Dimoulas A. Silicene and germanene: silicon and germanium in the ‘‘flatland”. Microelectron Eng 2015;131:68–78.
[51] Houssa M, Dimoulas A, Molle A. Silicene: a review of recent experimental and theoretical investigations. J Phys: Condens Matter 2015;27:253002.
[52] Friedlein R, Yamada-Takamura Y. Electronic properties of epitaxial silicene on diboride thin films. J Phys: Condens Matter 2015;27:203201.
[53] Grazianetti C, Cinquanta E, Molle A. Two-dimensional silicon: the advent of silicene. 2D Mater 2016;3:012001.
[54] Kaloni TP, Schreckenbach G, Freund MS, Schwingenschlögl U. Current developments in silicene and germanene; 2015. Available from: ArXiv:
151206300.
[55] Roome NJ, Carey JD. Beyond graphene: stable elemental monolayers of silicene and germanene. ACS Appl Mater Interfaces 2014;6:7743–50.
[56] Bocchetti V, Diep HT, Enriquez H, Oughaddou H, Kara A. Thermal stability of standalone silicene sheet. J Phys: Conf Ser 2014;491:012008.
[57] Berdiyorov GR, Peeters FM. Influence of vacancy defects on the thermal stability of silicene: a reactive molecular dynamics study. RSC Adv
2014;4:1133–7.
[58] Wang S. A comparative first-principles study of orbital hybridization in two-dimensional C, Si, and Ge. Phys Chem Chem Phys 2011;13:11929–38.
[59] Jose D, Datta A. Understanding of the buckling distortions in silicene. J Phys Chem C 2012;116:24639–48.
[60] Soto JR, Molina B, Castro JJ. Reexamination of the origin of the pseudo Jahn–Teller puckering instability in silicene. Phys Chem Chem Phys
2015;17:7624–8.
[61] Molina B, Soto JR, Castro JJ. Pseudo Jahn–Teller effect in the decasilanaphthalene molecule: towards the origin of the buckling in silicene. Chem Phys
2015;460:97–100.
[62] Changgu Lee XW, Kysar Jeffrey W, Hone James. Measurement of the elastic properties and intrinsic strength of monolayer graphene. Science
2008;321:385–8.
142 J. Zhao et al. / Progress in Materials Science 83 (2016) 24–151

[63] Zhang H, Wang R. The stability and the nonlinear elasticity of 2D hexagonal structures of Si and Ge from first-principles calculations. Physica B
2011;406:4080–4.
[64] Peng Q, Wen X, De S. Mechanical stabilities of silicene. RSC Adv 2013;3:13772.
[65] Xu P, Yu Z, Yang C, Lu P, Liu Y, Ye H, et al. Comparative study on the nonlinear properties of bilayer graphene and silicene under tension. Superlattices
Microstruct 2014;75:647–56.
[66] Botari T, Perim E, Autreto PA, van Duin AC, Paupitz R, Galvao DS. Mechanical properties and fracture dynamics of silicene membranes. Phys Chem
Chem Phys 2014;16:19417–23.
[67] Yang C, Yu Z, Lu P, Liu Y, Ye H, Gao T. Phonon instability and ideal strength of silicene under tension. Comput Mater Sci 2014;95:420–8.
[68] Wang B, Wu J, Gu X, Yin H, Wei Y, Yang R, et al. Stable planar single-layer hexagonal silicene under tensile strain and its anomalous Poisson’s ratio.
Appl Phys Lett 2014;104:081902.
[69] Zhao H. Strain and chirality effects on the mechanical and electronic properties of silicene and silicane under uniaxial tension. Phys Lett A
2012;376:3546–50.
[70] Roman RE, Cranford SW. Mechanical properties of silicene. Comput Mater Sci 2014;82:50–5.
[71] Pei Q-X, Sha Z-D, Zhang Y-Y, Zhang Y-W. Effects of temperature and strain rate on the mechanical properties of silicene. J Appl Phys 2014;115:023519.
[72] Gu X, Yang R. First-principles prediction of phononic thermal conductivity of silicene: a comparison with graphene. J Appl Phys 2015;117:025102.
[73] Nika D, Pokatilov E, Askerov A, Balandin A. Phonon thermal conduction in graphene: role of Umklapp and edge roughness scattering. Phys Rev B
2009;79:155413.
[74] Xu Y, Li Z, Duan W. Thermal and thermoelectric properties of graphene. Small 2014;10:2182–99.
[75] Xie H, Hu M, Bao H. Thermal conductivity of silicene from first-principles. Appl Phys Lett 2014;104:131906.
[76] Li X, Mullen J, Jin Z, Borysenko K, Buongiorno Nardelli M, Kim KW. Intrinsic electrical transport properties of monolayer silicene and MoS2 from first
principles. Phys Rev B 2013;87:115418.
[77] Huang L-F, Gong P-L, Zeng Z. Phonon properties, thermal expansion, and thermomechanics of silicene and germanene. Phys Rev B 2015;91:205433.
[78] Liu B, Reddy CD, Jiang J, Zhu H, Baimova JA, Dmitriev SV, et al. Thermal conductivity of silicene nanosheets and the effect of isotopic doping. J Phys D
Appl Phys 2014;47:165301.
[79] Ng TY, Yeo J, Liu Z. Molecular dynamics simulation of the thermal conductivity of shorts strips of graphene and silicene: a comparative study. Int J
Mech Mater Des 2013;9:105–14.
[80] Zhang X, Xie H, Hu M, Bao H, Yue S, Qin G, et al. Thermal conductivity of silicene calculated using an optimized Stillinger–Weber potential. Phys Rev B
2014;89:054310.
[81] Lepri S, Livi R, Politi A. Thermal conduction in classical low-dimensional lattices. Phys Rep 2003;377:1–80.
[82] Basile G, Bernardin C, Olla S. Momentum conserving model with anomalous thermal conductivity in low dimensional systems. Phys Rev Lett
2006;96:204303.
[83] Xu X, Pereira LF, Wang Y, Wu J, Zhang K, Zhao X, et al. Length-dependent thermal conductivity in suspended single-layer graphene. Nat Commun
2014;5:3689.
[84] Wang Z, Feng T, Ruan X. Thermal conductivity and spectral phonon properties of freestanding and supported silicene. J Appl Phys 2015;117:084317.
[85] Huang S, Kang W, Yang L. Electronic structure and quasiparticle bandgap of silicene structures. Appl Phys Lett 2013;102:133106.
[86] Matthes L, Gori P, Pulci O, Bechstedt F. Universal infrared absorbance of two-dimensional honeycomb group-IV crystals. Phys Rev B 2013;87:035438.
[87] Wei W, Dai Y, Huang B, Jacob T. Many-body effects in silicene, silicane, germanene and germanane. Phys Chem Chem Phys 2013;15:8789–94.
[88] Wei W, Jacob T. Strong many-body effects in silicene-based structures. Phys Rev B 2013;88:045203.
[89] Rojas-Cuervo AM, Fonseca-Romero KM, Rey-González RR. Anisotropic Dirac cones in monoatomic hexagonal lattices: a DFT study. Eur Phys J B
2014;87:67.
[90] Bechstedt F, Matthes L, Gori P, Pulci O. Infrared absorbance of silicene and germanene. Appl Phys Lett 2012;100:261906.
[91] Matthes L, Pulci O, Bechstedt F. Massive Dirac quasiparticles in the optical absorbance of graphene, silicene, germanene, and tinene. J Phys: Condens
Matter 2013;25:395305.
[92] Shao Z-G, Ye X-S, Yang L, Wang C-L. First-principles calculation of intrinsic carrier mobility of silicene. J Appl Phys 2013;114:093712.
[93] Mohan B, Kumar A, Ahluwalia PK. A first principle calculation of electronic and dielectric properties of electrically gated low-buckled mono and
bilayer silicene. Physica E 2013;53:233–9.
[94] Matthes L, Pulci O, Bechstedt F. Optical properties of two-dimensional honeycomb crystals graphene, silicene, germanene, and tinene from first
principles. New J Phys 2014;16:105007.
[95] Yang JY, Liu LH. Temperature-dependent dielectric functions in atomically thin graphene, silicene, and arsenene. Appl Phys Lett 2015;107:091902.
[96] Thouless D, Kohmoto M, Nightingale M, Den Nijs M. Quantized Hall conductance in a two-dimensional periodic potential. Phys Rev Lett 1982;49:405.
[97] Wen X-G. Topological orders and edge excitations in fractional quantum Hall states. Adv Phys 1995;44:405–73.
[98] Hasan MZ, Kane CL. Colloquium: topological insulators. Rev Mod Phys 2010;82:3045.
[99] Qi X-L, Zhang S-C. Topological insulators and superconductors. Rev Mod Phys 2011;83:1057.
[100] He K, Wang Y, Xue Q-K. Quantum anomalous Hall effect. Natl Sci Rev 2014;1:38–48.
[101] Fu L, Kane CL. Time reversal polarization and a Z2 adiabatic spin pump. Phys Rev B 2006;74:195312.
[102] Moore JE, Balents L. Topological invariants of time-reversal-invariant band structures. Phys Rev B 2007;75:121306.
[103] Xiao D, Chang M-C, Niu Q. Berry phase effects on electronic properties. Rev Mod Phys 2010;82:1959.
[104] Fukui T, Hatsugai Y. Quantum spin Hall effect in three dimensional materials: lattice computation of Z2 topological invariants and its application to Bi
and Sb. J Phys Soc Jpn 2007;76:053702.
[105] Xiao D, Yao Y, Feng W, Wen J, Zhu W, Chen X-Q, et al. Half-Heusler compounds as a new class of three-dimensional topological insulators. Phys Rev
Lett 2010;105:096404.
[106] Feng W, Xiao D, Ding J, Yao Y. Three-dimensional topological insulators in I–III–VI2 and II–IV–V2 chalcopyrite semiconductors. Phys Rev Lett
2011;106:016402.
[107] Prodan E. Robustness of the spin-Chern number. Phys Rev B 2009;80:125327.
[108] Prodan E. Non-commutative tools for topological insulators. New J Phys 2010;12:065003.
[109] Sheng D, Weng Z, Sheng L, Haldane F. Quantum spin-Hall effect and topologically invariant Chern numbers. Phys Rev Lett 2006;97:036808.
[110] Li H, Sheng L, Sheng D, Xing D. Chern number of thin films of the topological insulator Bi2Se3. Phys Rev B 2010;82:165104.
[111] Kane CL, Mele EJ. Quantum spin Hall effect in graphene. Phys Rev Lett 2005;95:226801.
[112] Kane CL, Mele EJ. Z2 topological order and the quantum spin Hall effect. Phys Rev Lett 2005;95:146802.
[113] Yao Y, Ye F, Qi X-L, Zhang S-C, Fang Z. Spin–orbit gap of graphene: first-principles calculations. Phys Rev B 2007;75:041401.
[114] Ezawa M. Valley-polarized metals and quantum anomalous Hall effect in silicene. Phys Rev Lett 2012;109:055502.
[115] Ezawa M. Photoinduced topological phase transition and a single Dirac-cone state in silicene. Phys Rev Lett 2013;110:026603.
[116] Pan H, Li Z, Liu C-C, Zhu G, Qiao Z, Yao Y. Valley-polarized quantum anomalous Hall effect in silicene. Phys Rev Lett 2014;112:106802.
[117] An X-T, Zhang Y-Y, Liu J-J, Li S-S. Quantum spin Hall effect induced by electric field in silicene. Appl Phys Lett 2013;102:043113.
[118] Geissler F, Budich JC, Trauzettel B. Group theoretical and topological analysis of the quantum spin Hall effect in silicene. New J Phys 2013;15:085030.
[119] Tabert CJ, Nicol EJ. Magneto-optical conductivity of silicene and other buckled honeycomb lattices. Phys Rev B 2013;88:085434.
[120] Tahir M, Manchon A, Sabeeh K, Schwingenschlögl U. Quantum spin/valley Hall effect and topological insulator phase transitions in silicene. Appl Phys
Lett 2013;102:162412.
[121] Tahir M, Schwingenschlögl U. Valley polarized quantum Hall effect and topological insulator phase transitions in silicene. Sci Rep 2013;3:1075.
J. Zhao et al. / Progress in Materials Science 83 (2016) 24–151 143

[122] Apalkov VM, Chakraborty T. Tunability of the fractional quantum Hall states in buckled Dirac materials. Phys Rev B 2014;90:245108.
[123] Shakouri K, Vasilopoulos P, Vargiamidis V, Peeters FM. Integer and half-integer quantum Hall effect in silicene: influence of an external electric field
and impurities. Phys Rev B 2014;90:235423.
[124] Pan H, Li X, Qiao Z, Liu C-C, Yao Y, Yang SA. Topological metallic phases in spin–orbit coupled bilayer systems. New J Phys 2014;16:123015.
[125] Ezawa M. Monolayer topological insulators: silicene, germanene and stanene. J Phys Soc Jpn 2015;84:121003.
[126] Ezawa M. Symmetry protected topological charge in symmetry broken phase: spin-Chern, spin-valley-Chern and mirror-Chern numbers. Phys Lett A
2014;378:1180–4.
[127] Yang Y, Xu Z, Sheng L, Wang B, Xing D, Sheng D. Time-reversal-symmetry-broken quantum spin Hall effect. Phys Rev Lett 2011;107:066602.
[128] Tsai WF, Huang CY, Chang TR, Lin H, Jeng HT, Bansil A. Gated silicene as a tunable source of nearly 100% spin-polarized electrons. Nat Commun
2013;4:1500.
[129] Varykhalov A, Sánchez-Barriga J, Shikin A, Biswas C, Vescovo E, Rybkin A, et al. Electronic and magnetic properties of quasifreestanding graphene on
Ni. Phys Rev Lett 2008;101:157601.
[130] Dedkov YS, Fonin M, Rüdiger U, Laubschat C. Rashba effect in the graphene/Ni(1 1 1) system. Phys Rev Lett 2008;100:107602.
[131] Ding J, Qiao Z, Feng W, Yao Y, Niu Q. Engineering quantum anomalous/valley Hall states in graphene via metal–atom adsorption: an ab-initio study.
Phys Rev B 2011;84:195444.
[132] Zhang XL, Liu LF, Liu WM. Quantum anomalous Hall effect and tunable topological states in 3d transition metals doped silicene. Sci Rep 2013;3:2908.
[133] Lin X, Ni J. Much stronger binding of metal adatoms to silicene than to graphene: a first-principles study. Phys Rev B 2012;86:075440.
[134] Sahin H, Peeters F. Adsorption of alkali, alkaline-earth, and 3d transition metal atoms on silicene. Phys Rev B 2013;87:085423.
[135] Ezawa M. Spin valleytronics in silicene: quantum spin Hall–quantum anomalous Hall insulators and single-valley semimetals. Phys Rev B
2013;87:155415.
[136] Pan H, Li X, Jiang H, Yao Y, Yang SA. Valley-polarized quantum anomalous Hall phase and disorder-induced valley-filtered chiral edge channels. Phys
Rev B 2015;91:045404.
[137] Haldane FDM. Model for a quantum Hall effect without Landau levels: condensed-matter realization of the ‘‘parity anomaly”. Phys Rev Lett
1988;61:2015.
[138] Wright AR. Realising Haldane’s vision for a Chern insulator in buckled lattices. Sci Rep 2013;3:2736.
[139] Kitagawa T, Oka T, Brataas A, Fu L, Demler E. Transport properties of nonequilibrium systems under the application of light: photoinduced quantum
Hall insulators without Landau levels. Phys Rev B 2011;84:235108.
[140] Eckardt A, Weiss C, Holthaus M. Superfluid-insulator transition in a periodically driven optical lattice. Phys Rev Lett 2005;95:260404.
[141] Inoue J-i, Tanaka A. Photoinduced transition between conventional and topological insulators in two-dimensional electronic systems. Phys Rev Lett
2010;105:017401.
[142] López A, Scholz A, Santos B, Schliemann J. Photoinduced pseudospin effects in silicene beyond the off-resonant condition. Phys Rev B
2015;91:125105.
[143] Shirley JH. Solution of the Schrödinger equation with a Hamiltonian periodic in time. Phys Rev 1965;138:B979.
[144] Novoselov KS, Geim AK, Morozov SV, Jiang D, Katsnelson MI, Grigorieva IV, et al. Two-dimensional gas of massless Dirac fermions in graphene. Nature
2005;438:197–200.
[145] Zhang Y, Tan Y-W, Stormer HL, Kim P. Experimental observation of the quantum Hall effect and Berry’s phase in graphene. Nature 2005;438:201–4.
[146] Ezawa M. Quantum Hall effects in silicene. J Phys Soc Jpn 2012;81:064705.
[147] Shakouri K, Vasilopoulos P, Vargiamidis V, Peeters FM. Spin- and valley-dependent magnetotransport in periodically modulated silicene. Phys Rev B
2014;90:125444.
[148] Shakouri K, Vasilopoulos P, Vargiamidis V, Hai GQ, Peeters FM. Spin- and valley-dependent commensurability oscillations and electric-field-induced
quantum Hall plateaux in periodically modulated silicene. Appl Phys Lett 2014;104:213109.
[149] Xi J, Long M, Tang L, Wang D, Shuai Z. First-principles prediction of charge mobility in carbon and organic nanomaterials. Nanoscale 2012;4:4348–69.
[150] Datta S. Electronic transport in mesoscopic systems. Cambridge: Cambridge University Press; 1997.
[151] Savini G, Ferrari A, Giustino F. First-principles prediction of doped graphane as a high-temperature electron–phonon superconductor. Phys Rev Lett
2010;105:037002.
[152] Yan J-A, Stein R, Schaefer DM, Wang X-Q, Chou MY. Electron–phonon coupling in two-dimensional silicene and germanene. Phys Rev B
2013;88:121403.
[153] Wang G. Do silicene nanoribbons have high carrier mobilities? Europhys Lett 2013;101:27005.
[154] Lee PA, Nagaosa N, Wen X-G. Doping a Mott insulator: physics of high-temperature superconductivity. Rev Mod Phys 2006;78:17.
[155] Armitage N, Fournier P, Greene R. Progress and perspectives on electron-doped cuprates. Rev Mod Phys 2010;82:2421.
[156] Stewart G. Superconductivity in iron compounds. Rev Mod Phys 2011;83:1589.
[157] Read N, Green D. Paired states of fermions in two dimensions with breaking of parity and time-reversal symmetries and the fractional quantum Hall
effect. Phys Rev B 2000;61:10267.
[158] Meng ZY, Kim YB, Kee H-Y. Odd-parity triplet superconducting phase in multiorbital materials with a strong spin–orbit coupling: application to
doped Sr2IrO4. Phys Rev Lett 2014;113:177003.
[159] Nayak C, Simon SH, Stern A, Freedman M, Sarma SD. Non-Abelian anyons and topological quantum computation. Rev Mod Phys 2008;80:1083.
[160] Alicea J. New directions in the pursuit of Majorana fermions in solid state systems. Rev Mod Phys 2012;75:076501.
[161] Wan W, Ge Y, Yang F, Yao Y. Phonon-mediated superconductivity in silicene predicted by first-principles density functional calculations. Europhys
Lett 2013;104:36001.
[162] Cheng YC, Zhu ZY, Schwingenschlögl U. Doped silicene: evidence of a wide stability range. Europhys Lett 2011;95:17005.
[163] Liu G, Wu MS, Ouyang CY, Xu B. Strain-induced semimetal-metal transition in silicene. Europhys Lett 2012;99:17010.
[164] Manes JL. Symmetry-based approach to electron–phonon interactions in graphene. Phys Rev B 2007;76:045430.
[165] Durajski AP, Szczȩśniak D, Szczȩśniak R. Study of the superconducting phase in silicene under biaxial tensile strain. Solid State Commun
2014;200:17–21.
[166] Zhang LD, Yang F, Yao Y. Possible electric-field-induced superconducting states in doped silicene. Sci Rep 2015;5:8203.
[167] Liu F, Liu C-C, Wu K, Yang F, Yao Y. d + id0 chiral superconductivity in bilayer silicene. Phys Rev Lett 2013;111:066804.
[168] Zhang L-D, Yang F, Yao Y. Itinerant ferromagnetism and p + ip0 superconductivity in doped bilayer silicene; 2015. Available from: ArXiv:150501971.
[169] Takimoto T, Hotta T, Ueda K. Strong-coupling theory of superconductivity in a degenerate Hubbard model. Phys Rev B 2004;69:104504.
[170] Kubo K. Pairing symmetry in a two-orbital Hubbard model on a square lattice. Phys Rev B 2007;75:224509.
[171] Graser S, Maier T, Hirschfeld P, Scalapino D. Near-degeneracy of several pairing channels in multiorbital models for the Fe pnictides. New J Phys
2009;11:025016.
[172] Nandkishore R, Levitov L, Chubukov A. Chiral superconductivity from repulsive interactions in doped graphene. Nat Phys 2012;8:158–63.
[173] McCann E, Fal’ko VI. Landau-level degeneracy and quantum Hall effect in a graphite bilayer. Phys Rev Lett 2006;96:086805.
[174] Partoens B, Peeters F. From graphene to graphite: electronic structure around the K point. Phys Rev B 2006;74:075404.
[175] Chen L, Liu C-C, Feng B, He X, Cheng P, Ding Z, et al. Evidence for Dirac fermions in a honeycomb lattice based on silicon. Phys Rev Lett
2012;109:056804.
[176] Enriquez H, Vizzini S, Kara A, Lalmi B, Oughaddou H. Silicene structures on silver surfaces. J Phys: Condens Matter 2012;24:314211.
[177] Lalmi B, Oughaddou H, Enriquez H, Kara A, Vizzini Sb, Ealet Bn, et al. Epitaxial growth of a silicene sheet. Appl Phys Lett 2010;97:223109.
144 J. Zhao et al. / Progress in Materials Science 83 (2016) 24–151

[178] Xun X, Jincheng Z, Yi D, Eilers S, Peleckis G, Waikong Y, et al. Epitaxial growth mechanism of silicene on Ag(1 1 1). In: 2014 International conference on
Nanoscience and Nanotechnology (ICONN). Adelaide, SA. IEEE; 2014. p. 28–30.
[179] Sone J, Yamagami T, Aoki Y, Nakatsuji K, Hirayama H. Epitaxial growth of silicene on ultra-thin Ag(1 1 1) films. New J Phys 2014;16:095004.
[180] Resta A, Leoni T, Barth C, Ranguis A, Becker C, Bruhn T, et al. Atomic structures of silicene layers grown on Ag(1 1 1): scanning tunneling microscopy
and noncontact atomic force microscopy observations. Sci Rep 2013;3:2399.
[181] Cinquanta E, Scalise E, Chiappe D, Grazianetti C, van den Broek B, Houssa M, et al. Getting through the nature of silicene: an sp2–sp3 two-dimensional
silicon nanosheet. J Phys Chem C 2013;117:16719–24.
[182] Zhuang J, Xu X, Du Y, Wu K, Chen L, Hao W, et al. Investigation of electron–phonon coupling in epitaxial silicene by in situ Raman spectroscopy. Phys
Rev B 2015;91:161409.
[183] Grazianetti C, Chiappe D, Cinquanta E, Fanciulli M, Molle A. Nucleation and temperature-driven phase transitions of silicene superstructures on Ag
(1 1 1). J Phys: Condens Matter 2015;27:255005.
p p
[184] Liu Z-L, Wang M-X, Liu C, Jia J-F, Vogt P, Quaresima C, et al. The fate of the 2 3  2 3R(30°) silicene phase on Ag(1 1 1). APL Mater 2014;2:092513.
[185] Kawahara K, Shirasawa T, Arafune R, Lin CL, Takahashi T, Kawai M, et al. Determination of atomic positions in silicene on Ag(1 1 1) by low-energy
electron diffraction. Surf Sci 2014;623:25–8.
[186] Acun A, Poelsema B, Zandvliet HJW, van Gastel R. The instability of silicene on Ag(1 1 1). Appl Phys Lett 2013;103:263119.
[187] Mannix AJ, Kiraly B, Fisher BL, Hersam MC, Guisinger NP. Silicon growth at the two-dimensional limit on Ag(1 1 1). ACS Nano 2014;8:7538–47.
[188] Prévot G, Bernard R, Cruguel H, Borensztein Y. Monitoring Si growth on Ag(1 1 1) with scanning tunneling microscopy reveals that silicene structure
involves silver atoms. Appl Phys Lett 2014;105:213106.
[189] Rahman MS, Nakagawa T, Mizuno S. Growth of Si on Ag(1 1 1) and determination of large commensurate unit cell of high-temperature phase. Jpn J
Appl Phys 2015;54:015502.
[190] Lin C-L, Arafune R, Kawahara K, Kanno M, Tsukahara N, Minamitani E, et al. Substrate-induced symmetry breaking in silicene. Phys Rev Lett
2013;110:076801.
[191] Guo Z-X, Furuya S, Iwata J-i, Oshiyama A. Absence of dirac electrons in silicene on Ag(1 1 1) surfaces. J Phys Soc Jpn 2013;82:063714.
[192] Tsoutsou D, Xenogiannopoulou E, Golias E, Tsipas P, Dimoulas A. Evidence for hybrid surface metallic band in (4  4) silicene on Ag(1 1 1). Appl Phys
Lett 2013;103:231604.
[193] Gori P, Pulci O, Ronci F, Colonna S, Bechstedt F. Origin of Dirac-cone-like features in silicon structures on Ag(1 1 1) and Ag(1 1 0). J Appl Phys
2013;114:113710.
[194] Cahangirov S, Audiffred M, Tang P, Iacomino A, Duan W, Merino G, et al. Electronic structure of silicene on Ag(1 1 1): strong hybridization effects. Phys
Rev B 2013;88:035432.
[195] Yuan Y, Quhe R, Zheng J, Wang Y, Ni Z, Shi J, et al. Strong band hybridization between silicene and Ag(1 1 1) substrate. Physica E 2014;58:38–42.
[196] Quhe R, Yuan Y, Zheng J, Wang Y, Ni Z, Shi J, et al. Does the Dirac cone exist in silicene on metal substrates? Sci Rep 2014;4:5476.
[197] Feng Y, Liu D, Feng B, Liu X, Zhao L, Xie Z, et al. Direct evidence of interaction-induced Dirac cones in monolayer silicene/Ag(1 1 1) system; 2015.
Available from: ArXiv:150306278.
[198] Avila J, De Padova P, Cho S, Colambo I, Lorcy S, Quaresima C, et al. Presence of gapped silicene-derived band in the prototypical (3  3) silicene phase
on silver (1 1 1) surfaces. J Phys: Condens Matter 2013;25:262001.
[199] Johnson NW, Vogt P, Resta A, De Padova P, Perez I, Muir D, et al. The metallic nature of epitaxial silicene monolayers on Ag(1 1 1). Adv Funct Mater
2014;24:5253–9.
[200] Liu Z-L, Wang M-X, Xu J-P, Ge J-F, Lay GL, Vogt P, et al. Various atomic structures of monolayer silicene fabricated on Ag(1 1 1). New J Phys
2014;16:075006.
[201] Arafune R, Lin C-L, Kawahara K, Tsukahara N, Minamitani E, Kim Y, et al. Structural transition of silicene on Ag(1 1 1). Surf Sci 2013;608:297–300.
[202] Kaltsas D, Tsetseris L, Dimoulas A. Structural evolution of single-layer films during deposition of silicon on silver: a first-principles study. J Phys:
Condens Matter 2012;24:442001.
[203] Guo Z-X, Furuya S, Iwata J-i, Oshiyama A. Absence and presence of Dirac electrons in silicene on substrates. Phys Rev B 2013;87:235435.
[204] N’Diaye AT, Bleikamp S, Feibelman PJ, Michely T. Two-dimensional Ir cluster lattice on a graphene Moire’ on Ir(1 1 1). Phys Rev Lett 2006;97:215501.
[205] Marchini S, Günther S, Wintterlin J. Scanning tunneling microscopy of graphene on Ru(0 0 0 1). Phys Rev B 2007;76:075429.
[206] Grazianetti C, Chiappe D, Cinquanta E, Tallarida G, Fanciulli M, Molle A. Exploring the morphological and electronic properties of silicene
superstructures. Appl Surf Sci 2014;291:109–12.
p p
[207] Tchalala MR, Enriquez H, Yildirim H, Kara A, Mayne AJ, Dujardin G, et al. Atomic and electronic structures of the ( 13  13)R13.9° of silicene sheet
on Ag(1 1 1). Appl Surf Sci 2014;303:61–6.
[208] Majzik Z, Rachid Tchalala M, Svec M, Hapala P, Enriquez H, Kara A, et al. Combined AFM and STM measurements of a silicene sheet grown on the Ag
(1 1 1) surface. J Phys: Condens Matter 2013;25:225301.
p p
[209] Enriquez H, Kara A, Mayne AJ, Dujardin G, Jamgotchian H, Aufray B, et al. Atomic structure of the (2 3  2 3)R30° of silicene on Ag(1 1 1) surface. J
Phys: Conf Ser 2014;491:012004.
p p
[210] Jamgotchian H, Ealet B, Colignon Y, Maradj H, Hoarau JY, Biberian JP, et al. A comprehensive study of the (2 3  2 3)R30° structure of silicene on Ag
(1 1 1). J Phys: Condens Matter 2015;27:395002.
p p
[211] Wang W, Olovsson W, Uhrberg RIG. Experimental and theoretical determination of r bands on (‘‘2 3  2 3”) silicene grown on Ag(1 1 1). Phys Rev B
2015;92:205427.
p p
[212] Jamgotchian H, Colignon Y, Maradj H, Hoarau J-Y, Biberian J-P, Aufray B. A comprehensive study of the (2 3  2 3)R30° structure of silicene on Ag
(1 1 1); 2015. Available from: ArXiv:14124902.
[213] Chen Jian, Feng Baojie, Cheng Peng, Qiu Jinglan, Chen Lan, Wu Kehui. Persistent Dirac fermion state on bulk-like Si(1 1 1) surface; 2015. Available
from: ArXiv:14057534.
[214] Vogt P, Capiod P, Berthe M, Resta A, De Padova P, Bruhn T, et al. Synthesis and electrical conductivity of multilayer silicene. Appl Phys Lett
2014;104:021602.
[215] Salomon E, El Ajjouri R, Le Lay G, Angot T. Growth and structural properties of silicene at multilayer coverage. J Phys: Condens Matter
2014;26:185003.
[216] Lagarde P, Chorro M, Roy D, Trcera N. Study by EXAFS of the local structure around Si on silicene deposited on Ag(1 1 0) and Ag(1 1 1) surfaces. J Phys:
Condens Matter 2016;28:075002.
[217] Fu H, Chen L, Chen J, Qiu J, Ding Z, Zhang J, et al. Multilayered silicene: the bottom-up approach for a weakly relaxed Si(1 1 1) with Dirac surface states.
Nanoscale 2015;7:15880–5.
[218] Chen L, Li H, Feng B, Ding Z, Qiu J, Cheng P, et al. Spontaneous symmetry breaking and dynamic phase transition in monolayer silicene. Phys Rev Lett
2013;110:085504.
p p
[219] Cahangirov S, Özçelik VO, Xian L, Avila J, Cho S, Asensio MC, et al. Atomic structure of the 3  3 phase of silicene on Ag(1 1 1). Phys Rev B
2014;90:035448.
[220] Guo Z-X, Oshiyama A. Crossover between silicene and ultra-thin Si atomic layers on Ag(1 1 1) surfaces. New J Phys 2015;17:045028.
[221] Wang Y-P, Cheng H-P. Absence of a Dirac cone in silicene on Ag(1 1 1): first-principles density functional calculations with a modified effective band
structure technique. Phys Rev B 2013;87:245430.
[222] Chen MX, Weinert M. Revealing the substrate origin of the linear dispersion of silicene/Ag(1 1 1). Nano Lett 2014;14:5189–93.
[223] Ishida H, Hamamoto Y, Morikawa Y, Minamitani E, Arafune R, Takagi N. Electronic structure of the 4  4 silicene monolayer on semi-infinite Ag(1 1 1).
New J Phys 2015;17:015013.
J. Zhao et al. / Progress in Materials Science 83 (2016) 24–151 145

[224] Mahatha SK, Moras P, Bellini V, Sheverdyaeva PM, Struzzi C, Petaccia L, et al. Silicene on Ag(1 1 1): a honeycomb lattice without Dirac bands. Phys Rev
B 2014;89:201416.
[225] Pflugradt P, Matthes L, Bechstedt F. Silicene-derived phases on Ag(1 1 1) substrate versus coverage: ab initio studies. Phys Rev B 2014;89:035403.
[226] Houssa M, van den Broek B, Scalise E, Ealet B, Pourtois G, Chiappe D, et al. Theoretical aspects of graphene-like group IV semiconductors. Appl Surf Sci
2014;291:98–103.
[227] Kaltsas D, Tsetseris L, Dimoulas A. Silicene on metal substrates: a first-principles study on the emergence of a hierarchy of honeycomb structures.
Appl Surf Sci 2014;291:93–7.
[228] Scalise E, Cinquanta E, Houssa M, van den Broek B, Chiappe D, Grazianetti C, et al. Vibrational properties of epitaxial silicene layers on (1 1 1) Ag. Appl
Surf Sci 2014;291:113–7.
[229] Stephan R, Hanf MC, Sonnet P. Spatial analysis of interactions at the silicene/Ag interface: first principles study. J Phys: Condens Matter
2015;27:015002.
[230] Houssa M, van den Broek B, Scalise E, Pourtois G, Afanas’ev V, Stesmans A. Theoretical study of silicene and germanene. ECS Trans 2013;53:51–62.
[231] Guo Z-X, Oshiyama A. Structural tristability and deep Dirac states in bilayer silicene on Ag(1 1 1) surfaces. Phys Rev B 2014;89:155418.
[232] Liu H, Gao J, Zhao J. Silicene on substrates: interaction mechanism and growth behavior. J Phys: Conf Ser 2014;491:012007.
[233] Zhu X, Zeng XC. Structures and stabilities of small silicon clusters: ab initio molecular-orbital calculations of Si7–Si11. J Chem Phys 2003;118:3558–70.
[234] Zhu XL, Zeng XC, Lei YA, Pan B. Structures and stability of medium silicon clusters. II. Ab initio molecular orbital calculations of Si12–Si20. J Chem Phys
2004;120:8985–95.
[235] Yoo S, Zeng XC. Structures and relative stability of medium-sized silicon clusters. IV. Motif-based low-lying clusters Si21–Si30. J Chem Phys
2006;124:054304.
[236] Shu H, Cao D, Liang P, Wang X, Chen X, Lu W. Two-dimensional silicene nucleation on a Ag(1 1 1) surface: structural evolution and the role of surface
diffusion. Phys Chem Chem Phys 2014;16:304–10.
[237] Rahman MS, Nakagawa T, Mizuno S. Study the surface structure evolution of Si-adsorption on Ag(1 1 1) by LEED-AES. In: 2014 International
conference on informatics, electronics & vision (ICIEV). Dhaka. p. 1–4.
[238] Feng B, Li H, Liu C-C, Shao T-N, Cheng P, Yao Y, et al. Observation of Dirac cone warping and chirality effects in silicene. ACS Nano 2013;7:9049–54.
[239] Molle A, Chiappe D, Cinquanta E, Grazianetti C, Fanciulli M, Scalise E, et al. Structural and chemical stabilization of the epitaxial silicene. ECS Trans
2013;58:217–27.
[240] Fukaya Y, Mochizuki I, Maekawa M, Wada K, Hyodo T, Matsuda I, et al. Structure of silicene on a Ag(1 1 1) surface studied by reflection high-energy
positron diffraction. Phys Rev B 2013;88:205413.
[241] Cinquanta E, Fratesi G, dal Conte S, Grazianetti C, Scotognella F, Stagira S, et al. Optical response and ultrafast carrier dynamics of the silicene–silver
interface. Phys Rev B 2015;92:165427.
[242] Scalise E, Houssa M, Pourtois G, van den Broek B, Afanas’ev V, Stesmans A. Vibrational properties of silicene and germanene. Nano Res 2012;6:19–28.
[243] Fleurence A, Yoshida Y, Lee CC, Ozaki T, Yamada-Takamura Y, Hasegawa Y. Microscopic origin of the p states in epitaxial silicene. Appl Phys Lett
2014;104:021605.
[244] Friedlein R, Fleurence A, Aoyagi K, de Jong MP, Van Bui H, Wiggers FB, et al. Core level excitations – a fingerprint of structural and electronic properties
of epitaxial silicene. J Chem Phys 2014;140:184704.
[245] Aizawa T, Suehara S, Otani S. Phonon dispersion of silicene on ZrB2(0 0 0 1). J Phys: Condens Matter 2015;27:305002.
[246] Wei W, Dai Y, Huang B, Whangbo M-H, Jacob T. Loss of linear band dispersion and trigonal structure in silicene on Ir(1 1 1). J Phys Chem Lett
2015;6:1065–70.
[247] Švec M, Hapala P, Ondráček M, Merino P, Blanco-Rey M, Mutombo P, et al. Silicene versus two-dimensional ordered silicide: atomic and electronic
p p
structure of Si-( 19  19)R23.4°/Pt(1 1 1). Phys Rev B 2014;89:201412.
[248] Lee C-C, Fleurence A, Friedlein R, Yamada-Takamura Y, Ozaki T. Avoiding critical-point phonon instabilities in two-dimensional materials: the origin
of the stripe formation in epitaxial silicene. Phys Rev B 2014;90:241402.
[249] Van Bui H, Wiggers FB, Friedlein R, Yamada-Takamura Y, Kovalgin AY, de Jong MP. On the feasibility of silicene encapsulation by AlN deposited using
an atomic layer deposition process. J Chem Phys 2015;142:064702.
[250] Lee C-C, Fleurence A, Friedlein R, Yamada-Takamura Y, Ozaki T. First-principles study on competing phases of silicene: effect of substrate and strain.
Phys Rev B 2013;88:165404.
[251] Lee C-C, Fleurence A, Yamada-Takamura Y, Ozaki T, Friedlein R. Band structure of silicene on zirconium diboride (0 0 0 1) thin-film surface:
convergence of experiment and calculations in the one-Si-atom Brillouin zone. Phys Rev B 2014;90:075422.
[252] Podsiadly-Paszkowska A, Krawiec M. Dirac fermions in silicene on Pb(1 1 1) surface. Phys Chem Chem Phys 2015;17:2246–51.
[253] Pflugradt P, Matthes L, Bechstedt F. Silicene on metal and metallized surfaces: ab initio studies. New J Phys 2014;16:075004.
[254] Morishita T, Spencer MJS, Kawamoto S, Snook IK. A new surface and structure for silicene: polygonal silicene formation on the Al(1 1 1) surface. J Phys
Chem C 2013;117:22142–8.
[255] Kanno M, Arafune R, Liang Lin C, Minamitani E, Kawai M, Takagi N. Electronic decoupling by h-BN layer between silicene and Cu(1 1 1): a DFT-based
analysis. New J Phys 2014;16:105019.
[256] Derivaz M, Dentel D, Stephan R, Hanf M-C, Mehdaoui A, Sonnet P, et al. Continuous germanene layer on Al(1 1 1). Nano Lett 2015;15:2510–6.
[257] Stephan R, Hanf MC, Derivaz M, Dentel D, Asensio MC, Avila J, et al. Germanene on Al(1 1 1): interface electronic states and charge transfer. J Phys
Chem C 2016;120:1580–5.
[258] Bhattacharya A, Bhattacharya S, Das GP. Exploring semiconductor substrates for silicene epitaxy. Appl Phys Lett 2013;103:123113.
[259] Neek-Amal M, Sadeghi A, Berdiyorov GR, Peeters FM. Realization of free-standing silicene using bilayer graphene. Appl Phys Lett 2013;103:261904.
[260] Kokott S, Pflugradt P, Matthes L, Bechstedt F. Nonmetallic substrates for growth of silicene: an ab initio prediction. J Phys: Condens Matter
2014;26:185002.
[261] Jia X, Hofmann M, Meunier V, Sumpter BG, Campos-Delgado J, Romo-Herrera JM, et al. Controlled formation of sharp zigzag and armchair edges in
graphitic nanoribbons. Science 2009;323:1701–5.
[262] Han MY, Özyilmaz B, Zhang Y, Kim P. Energy band-gap engineering of graphene nanoribbons. Phys Rev Lett 2007;98:206805.
[263] An X-T, Zhang Y-Y, Liu J-J, Li S-S. Interplay between edge and bulk states in silicene nanoribbon. Appl Phys Lett 2013;102:213115.
[264] Kang J, Wu F, Li J. Symmetry-dependent transport properties and magnetoresistance in zigzag silicene nanoribbons. Appl Phys Lett 2012;100:233122.
[265] Ezawa M. Quantized conductance and field-effect topological quantum transistor in silicene nanoribbons. Appl Phys Lett 2013;102:172103.
[266] Leandri C, Lay GL, Aufray B, Girardeaux C, Avila J, Dávila ME, et al. Self-aligned silicon quantum wires on Ag(1 1 0). Surf Sci 2005;574:L9–L15.
[267] Kara A, Léandri C, Dávila ME, De Padova P, Ealet B, Oughaddou H, et al. Physics of silicene stripes. J Supercond Nov Magn 2009;22:259–63.
[268] Colonna S, Serrano G, Gori P, Cricenti A, Ronci F. Systematic STM and LEED investigation of the Si/Ag(1 1 0) surface. J Phys: Condens Matter
2013;25:315301.
[269] Bernard R, Leoni T, Wilson A, Lelaidier T, Sahaf H, Moyen E, et al. Growth of Si ultrathin films on silver surfaces: evidence of an Ag(1 1 0) reconstruction
induced by Si. Phys Rev B 2013;88:121411(R).
[270] Tchalala MR, Enriquez H, Mayne AJ, Kara A, Dujardin G, Ali MA, et al. Atomic structure of silicene nanoribbons on Ag(1 1 0). J Phys: Conf Ser
2014;491:012002.
[271] Feng B, Li H, Meng S, Chen L, Wu K. Structure and quantum well states in silicene nanoribbons on Ag(1 1 0). Surf Sci 2016;645:74–9.
[272] Ronci F, Colonna S, Cricenti A, De Padova P, Ottaviani C, Quaresima C, et al. Low temperature STM/STS study of silicon nanowires grown on the Ag
(1 1 0) surface. Phys Status Solidi C 2010;7:2716–9.
146 J. Zhao et al. / Progress in Materials Science 83 (2016) 24–151

[273] Le Lay G, Aufray B, Léandri C, Oughaddou H, Biberian JP, De Padova P, et al. Physics and chemistry of silicene nano-ribbons. Appl Surf Sci
2009;256:524–9.
[274] De Padova P, Kubo O, Olivieri B, Quaresima C, Nakayama T, Aono M, et al. Multilayer silicene nanoribbons. Nano Lett 2012;12:5500–3.
[275] Kara A, Vizzini S, Leandri C, Ealet B, Oughaddou H, Aufray B, et al. Silicon nano-ribbons on Ag(1 1 0): a computational investigation. J Phys: Condens
Matter 2010;22:045004.
[276] Lian C, Ni J. The structural and electronic properties of silicon nanoribbons on Ag(1 1 0): a first principles study. Physica B 2012;407:4695–9.
[277] Speiser E, Buick B, Esser N, Richter W, Colonna S, Cricenti A, et al. Raman spectroscopy study of silicon nanoribbons on Ag(1 1 0). Appl Phys Lett
2014;104:161612.
[278] Ronci F, Serrano G, Gori P, Cricenti A, Colonna S. Silicon-induced faceting at the Ag(1 1 0) surface. Phys Rev B 2014;89:115437.
[279] Hogan C, Colonna S, Flammini R, Cricenti A, Ronci F. Structure and stability of Si/Ag(1 1 0) nanoribbons. Phys Rev B 2015;92:115439.
[280] Bai J, Tanaka H, Zeng XC. Graphene-like bilayer hexagonal silicon polymorph. Nano Res 2010;3:694–700.
[281] Morishita T, Russo SP, Snook IK, Spencer MJS, Nishio K, Mikami M. First-principles study of structural and electronic properties of ultrathin silicon
nanosheets. Phys Rev B 2010;82:045419.
[282] Pan L, Liu HJ, Wen YW, Tan XJ, Lv HY, Shi J, et al. First-principles study of monolayer and bilayer honeycomb structures of group-IV elements and their
binary compounds. Phys Lett A 2011;375:614–9.
[283] Padilha JE, Pontes RB. Free-standing bilayer silicene: the effect of stacking order on the structural, electronic, and transport properties. J Phys Chem C
2015;119:3818–25.
[284] Naji S, Khalil B, Labrim H, Bhihi M, Belhaj A, Benyoussef A, et al. Interdistance effects on flat and buckled silicene like-bilayers. J Phys: Conf Ser
2014;491:012006.
[285] Fu H, Zhang J, Ding Z, Li H, Meng S. Stacking-dependent electronic structure of bilayer silicene. Appl Phys Lett 2014;104:131904.
[286] Sakai Y, Oshiyama A. Structural stability and energy-gap modulation through atomic protrusion in freestanding bilayer silicene. Phys Rev B
2015;91:201405.
[287] Spencer MJ, Morishita T, Snook IK. Reconstruction and electronic properties of silicon nanosheets as a function of thickness. Nanoscale
2012;4:2906–13.
[288] Kamal C, Chakrabarti A, Banerjee A, Deb SK. Silicene beyond mono-layers – different stacking configurations and their properties. J Phys: Condens
Matter 2013;25:085508.
[289] Guo Z-X, Zhang Y-Y, Xiang H, Gong X-G, Oshiyama A. Structural evolution and optoelectronic applications of multilayer silicene. Phys Rev B
2015;92:201413.
[290] Johnson NW, Muir D, Kurmaev EZ, Moewes A. Stability and electronic characteristics of epitaxial silicene multilayers on Ag(1 1 1). Adv Funct Mater
2015;25:4083–90.
[291] De Padova P, Ottaviani C, Quaresima C, Olivieri B, Imperatori P, Salomon E, et al. 24 h stability of thick multilayer silicene in air. 2D Mater
2014;1:021003.
[292] Pflugradt P, Matthes L, Bechstedt F. Unexpected symmetry and AA stacking of bilayer silicene on Ag(1 1 1). Phys Rev B 2014;89:205428.
p p
[293] Borensztein Y, Curcella A, Royer S, Prévot G. Silicene multilayers on Ag(1 1 1) display a cubic diamond like structure and a 3  3 reconstruction
induced by surfactant Ag atoms. Phys Rev B 2015;92:155407.
[294] Shirai T, Shirasawa T, Hirahara T, Fukui N, Takahashi T, Hasegawa S. Structure determination of multilayer silicene grown on Ag(1 1 1) films by
electron diffraction: evidence for Ag segregation at the surface. Phys Rev B 2014;89:241403.
[295] Chen J, Feng B, Cheng P, Qiu J, Chen L, Wu K. unpublished.
[296] Matsuda I, Morikawa H, Liu C, Ohuchi S, Hasegawa S, Okuda T, et al. Electronic evidence of asymmetry in the mathrm Si(1 1 1) structure. Phys Rev B
2003;68:085407.
[297] Yaokawa R, Ohsuna T, Morishita T, Hayasaka Y, Spencer MJ, Nakano H. Monolayer-to-bilayer transformation of silicenes and their structural analysis.
Nat Commun 2016;7:10657.
[298] Houssa M, Scalise E, Sankaran K, Pourtois G, Afanas’ev VV, Stesmans A. Electronic properties of hydrogenated silicene and germanene. Appl Phys Lett
2011;98:223107.
[299] Sofo JO, Chaudhari AS, Barber GD. Graphane: a two-dimensional hydrocarbon. Phys Rev B 2007;75:153401.
[300] Lebègue S, Klintenberg M, Eriksson O, Katsnelson MI. Accurate electronic band gap of pure and functionalized graphane from GW calculations. Phys
Rev B 2009;79:245117.
[301] Zhang P, Li XD, Hu CH, Wu SQ, Zhu ZZ. First-principles studies of the hydrogenation effects in silicene sheets. Phys Lett A 2012;376:1230–3.
[302] Zólyomi V, Wallbank JR, Fal’ko VI. Silicane and germanane: tight-binding and first-principles studies. 2D Mater 2014;1:011005.
[303] Peng Q, De S. Elastic limit of silicane. Nanoscale 2014;6:12071–9.
[304] Ding Y, Wang Y. Electronic structures of silicene fluoride and hydride. Appl Phys Lett 2012;100:083102.
[305] Wu W, Ao Z, Wang T, Li C, Li S. Electric field induced hydrogenation of silicene. Phys Chem Chem Phys 2014;16:16588–94.
[306] Nguyen M-T, Phong PN, Tuyen ND. Hydrogenated graphene and hydrogenated silicene: computational insights. ChemPhysChem 2015;16:1733–8.
[307] Zhang C-w, Yan S-s. First-principles study of ferromagnetism in two-dimensional silicene with hydrogenation. J Phys Chem C 2012;116:4163–6.
[308] Osborn TH, Farajian AA, Pupysheva OV, Aga RS, Lew Yan Voon LC. Ab initio simulations of silicene hydrogenation. Chem Phys Lett 2011;511:101–5.
[309] Cao G, Zhang Y, Cao J. Strain and chemical function decoration induced quantum spin Hall effect in 2D silicene and Sn film. Phys Lett A
2015;379:1475–9.
[310] Guzman-Verri GG, Lew Yan Voon LC. Band structure of hydrogenated Si nanosheets and nanotubes. J Phys: Condens Matter 2011;23:145502.
[311] van den Broek B, Houssa M, Scalise E, Pourtois G, Afanas‘ev VV, Stesmans A. First-principles electronic functionalization of silicene and germanene by
adatom chemisorption. Appl Surf Sci 2014;291:104–8.
[312] Garcia JC, de Lima DB, Assali LVC, Justo JoF. Group IV graphene- and graphane-like nanosheets. J Phys Chem C 2011;115:13242–6.
[313] Wang XQ, Li HD, Wang JT. Induced ferromagnetism in one-side semihydrogenated silicene and germanene. Phys Chem Chem Phys 2012;14:3031–6.
[314] Zheng F-b, Zhang C-w. The electronic and magnetic properties of functionalized silicene: a first-principles study. Nanoscale Res Lett 2012;7:422.
[315] Pi X, Ni Z, Liu Y, Ruan Z, Xu M, Yang D. Density functional theory study on boron- and phosphorus-doped hydrogen-passivated silicene. Phys Chem
Chem Phys 2015;17:4146–51.
[316] Huang B, Deng H-X, Lee H, Yoon M, Sumpter BG, Liu F, et al. Exceptional optoelectronic properties of hydrogenated bilayer silicene. Phys Rev X
2014;4:021029.
[317] Rupp CJ, Chakraborty S, Ahuja R, Baierle RJ. The effect of impurities in ultra-thin hydrogenated silicene and germanene: a first principles study. Phys
Chem Chem Phys 2015;17:22210–6.
[318] Rupp CJ, Chakraborty S, Anversa J, Baierle RJ, Ahuja R. Rationalizing the hydrogen and oxygen evolution reaction activity of two-dimensional
hydrogenated silicene and germanene. ACS Appl Mater Interfaces 2016;8:1536–44.
[319] Liu Y, Shu H, Liang P, Cao D, Chen X, Lu W. Structural, electronic, and optical properties of hydrogenated few-layer silicene: size and stacking effects. J
Appl Phys 2013;114:094308.
[320] Wang R, Wang S, Wu X. The formation and electronic properties of hydrogenated bilayer silicene from first-principles. J Appl Phys 2014;116:024303.
[321] Lin S-Y, Chang S-L, Thuy Tran NT, Yang P-H, Lin M-F. H–Si bonding-induced unusual electronic properties of silicene: a method to identify hydrogen
concentration. Phys Chem Chem Phys 2015;17:26443–50.
[322] Qiu J, Fu H, Xu Y, Zhou Q, Meng S, Li H, et al. From silicene to half-silicane by hydrogenation. ACS Nano 2015;9:11192.
[323] Gao N, Zheng WT, Jiang Q. Density functional theory calculations for two-dimensional silicene with halogen functionalization. Phys Chem Chem Phys
2012;14:257–61.
J. Zhao et al. / Progress in Materials Science 83 (2016) 24–151 147

[324] Denis PA. Stacked functionalized silicene: a powerful system to adjust the electronic structure of silicene. Phys Chem Chem Phys 2015;17:5393–402.
[325] Wang X, Liu H, Tu S-T. First-principles study of half-fluorinated silicene sheets. RSC Adv 2015;5:6238–45.
[326] Huang B, Xiang HJ, Wei S-H. Chemical functionalization of silicene: spontaneous structural transition and exotic electronic properties. Phys Rev Lett
2013;111:145502.
[327] Noor-A-Alam M, Kim HJ, Shin Y-H. Hydrogen and fluorine co-decorated silicene: a first principles study of piezoelectric properties. J Appl Phys
2015;117:224304.
[328] Wang N, Guo H, Liu Y-j, Zhao J-x, Cai Q-h, Wang X-z. Asymmetric functionalization as a promising route to open the band gap of silicene: a theoretical
prediction. Physica E 2015;73:21–6.
[329] Zhao T, Zhang S, Wang Q, Kawazoe Y, Jena P. Tuning electronic and magnetic properties of silicene with magnetic superhalogens. Phys Chem Chem
Phys 2014;16:22979–86.
[330] De Padova P, Quaresima C, Olivieri B, Perfetti P, Le Lay G. Strong resistance of silicene nanoribbons towards oxidation. J Phys D Appl Phys
2011;44:312001.
[331] Molle A, Grazianetti C, Chiappe D, Cinquanta E, Cianci E, Tallarida G, et al. Hindering the oxidation of silicene with non-reactive encapsulation. Adv
Funct Mater 2013;23:4340–4.
[332] Friedlein R, Van Bui H, Wiggers FB, Yamada-Takamura Y, Kovalgin AY, de Jong MP. Interaction of epitaxial silicene with overlayers formed by
exposure to Al atoms and O2 molecules. J Chem Phys 2014;140:204705.
[333] Liu G, Lei XL, Wu MS, Xu B, Ouyang CY. Is silicene stable in O2?—first-principles study of O2 dissociation and O2-dissociation-induced oxygen atoms
adsorption on free-standing silicene. Europhys Lett 2014;106:47001.
[334] Liu G, Lei XL, Wu MS, Xu B, Ouyang CY. Comparison of the stability of free-standing silicene and hydrogenated silicene in oxygen: a first principles
investigation. J Phys: Condens Matter 2014;26:355007.
[335] Morishita T, Spencer MJ. How silicene on Ag(1 1 1) oxidizes: microscopic mechanism of the reaction of O2 with silicene. Sci Rep 2015;5:17570.
[336] Xu X, Zhuang J, Du Y, Feng H, Zhang N, Liu C, et al. Effects of oxygen adsorption on the surface state of epitaxial silicene on Ag(1 1 1). Sci Rep
2014;4:7543.
[337] Osborn TH, Farajian AA. Stability of lithiated silicene from first principles. J Phys Chem C 2012;116:22916–20.
[338] Kaloni TP, Schwingenschlögl U. Effects of heavy metal adsorption on silicene. Phys Status Solidi RRL 2014;8:685–7.
[339] Gurel HH, Ozcelik VO, Ciraci S. Effects of charging and perpendicular electric field on the properties of silicene and germanene. J Phys: Condens Matter
2013;25:305007.
[340] Ersan F, Arslanalp Ö, Gökoğlu G, Aktürk E. Effects of silver adatoms on the electronic structure of silicene. Appl Surf Sci 2014;311:9–13.
[341] Sivek J, Sahin H, Partoens B, Peeters F. Adsorption and absorption of boron, nitrogen, aluminum, and phosphorus on silicene: stability and electronic
and phonon properties. Phys Rev B 2013;87:085444.
[342] Kaloni TP, Schreckenbach G, Freund MS. Large enhancement and tunable band gap in silicene by small organic molecule adsorption. J Phys Chem C
2014;118:23361–7.
[343] Kaloni TP, Gangopadhyay S, Singh N, Jones B, Schwingenschlögl U. Electronic properties of Mn-decorated silicene on hexagonal boron nitride. Phys
Rev B 2013;88:235418.
[344] Kaloni TP, Singh N, Schwingenschlögl U. Prediction of a quantum anomalous Hall state in Co-decorated silicene. Phys Rev B 2014;89:035409.
[345] Zhang J, Zhao B, Yang Z. Abundant topological states in silicene with transition metal adatoms. Phys Rev B 2013;88:165422.
[346] Zheng R, Chen Y, Ni J. Highly tunable magnetism in silicene doped with Cr and Fe atoms under isotropic and uniaxial tensile strain. Appl Phys Lett
2015;107:263104.
[347] Han N, Liu H, Zhao J. Novel magnetic monolayers of transition metal silicide. J Supercond Nov Magn 2015;28:1755–8.
[348] Dzade NY, Obodo KO, Adjokatse SK, Ashu AC, Amankwah E, Atiso CD, et al. Silicene and transition metal based materials: prediction of a two-
dimensional piezomagnet. J Phys: Condens Matter 2010;22:375502.
[349] Sun X, Wang L, Lin H, Hou T, Li Y. Induce magnetism into silicene by embedding transition-metal atoms. Appl Phys Lett 2015;106:222401.
[350] Friedlein R, Fleurence A, Sadowski JT, Yamada-Takamura Y. Tuning of silicene–substrate interactions with potassium adsorption. Appl Phys Lett
2013;102:221603.
[351] Noguchi E, Sugawara K, Yaokawa R, Hitosugi T, Nakano H, Takahashi T. Direct observation of Dirac cone in multilayer silicene intercalation compound
CaSi2. Adv Mater 2015;27:856–60.
[352] Dutta S, Wakabayashi K. Momentum shift of Dirac cones in the silicene-intercalated compound CaSi2. Phys Rev B 2015;91:201410.
[353] Cahangirov S, Topsakal M, Ciraci S. Armchair nanoribbons of silicon and germanium honeycomb structures. Phys Rev B 2010;81:195120.
[354] Li R, Zhou J, Han Y, Dong J, Kawazoe Y. A new (2  1) reconstructed edge structure of zigzag Si nanoribbon: first principles study. J Chem Phys
2013;139:104703.
[355] Ding Y, Wang Y. Electronic structures of reconstructed zigzag silicene nanoribbons. Appl Phys Lett 2014;104:083111.
[356] Song Y-L, Zhang Y, Zhang J-M, Lu D-B. Effects of the edge shape and the width on the structural and electronic properties of silicene nanoribbons. Appl
Surf Sci 2010;256:6313–7.
[357] Song Y-L, Zhang S, Lu D-B, H-r Xu, Wang Z, Zhang Y, et al. Band-gap modulations of armchair silicene nanoribbons by transverse electric fields. Eur
Phys J B 2013;86:488.
[358] Liang Y, Wang V, Mizuseki H, Kawazoe Y. Band gap engineering of silicene zigzag nanoribbons with perpendicular electric fields: a theoretical study. J
Phys: Condens Matter 2012;24:455302.
[359] Ezawa M, Nagaosa N. Interference of topologically protected edge states in silicene nanoribbons. Phys Rev B 2013;88:121401.
[360] Kikutake K, Ezawa M, Nagaosa N. Edge states in silicene nanodisks. Phys Rev B 2013;88:115432.
[361] Zhao YC, Ni J. Spin-semiconducting properties in silicene nanoribbons. Phys Chem Chem Phys 2014;16:15477–82.
[362] Ding Y, Wang Y. Electronic structures of zigzag silicene nanoribbons with asymmetric sp2–sp3 edges. Appl Phys Lett 2013;102:143115.
[363] Song Y-L, Zhang Y, Zhang J-M, Lu D-B, Xu K-W. Tuning the electronic and magnetic properties of the Si nanoribbons through dangling bond. Physica B
2011;406:699–704.
[364] Fang DQ, Zhang Y, Zhang SL. Silicane nanoribbons: electronic structure and electric field modulation. New J Phys 2014;16:115006.
[365] Luan H-x, Zhang C-w, Li F, Wang Pj. Electronic and magnetic properties of silicene nanoflakes by first-principles calculations. Phys Lett A
2013;377:2792–5.
[366] Wang Y, Zheng J, Ni Z, Fei R, Liu Q, Quhe R, et al. Half-metallic silicene and germanene nanoribbons: towards high-performance spintronics device.
NANO 2012;07:1250037.
[367] Li G, Tan J, Liu X, Wang X, Li F, Zhao M. Manifold electronic structure transition of hybrid silicane–silicene nanoribbons. Chem Phys Lett 2014;595–
596:20–4.
[368] Zhang D, Long M, Zhang X, Cao C, Xu H, Li M, et al. Bipolar spin-filtering, rectifying and giant magnetoresistance effects in zigzag silicene nanoribbons
with asymmetric edge hydrogenation. Chem Phys Lett 2014;616–617:178–83.
[369] Jiang QG, Zhang JF, Ao ZM, Wu YP. Density functional theory study on the electronic properties and stability of silicene/silicane nanoribbons. J Mater
Chem C 2015;3:3954–9.
[370] Mohan B, Pooja, Kumar A, Ahluwalia PK. Shape and edge dependent electronic and magnetic properties of silicene nano-flakes. AIP Conf Proc, 1665. p.
140054.
[371] Chowdhury S, Nath P, Jana D. Shape dependent magnetic and optical properties in silicene nanodisks: a first principles study. J Phys Chem Solids
2015;83:32–9.
148 J. Zhao et al. / Progress in Materials Science 83 (2016) 24–151

[372] Jiang QG, Zhang JF, Ao ZM, Wu YP. First principles study on the electronic structure and interface stability of hybrid silicene/fluorosilicene
nanoribbons. Sci Rep 2015;5:15734.
[373] Abdelsalam H, Espinosa-Ortega T, Lukyanchuk I. Tuning of zero energy states in quantum dots of Silicene and bilayer graphene by electric field.
Superlattices Microstruct 2015;87:137–42.
[374] Banhart F, Kotakoski J, Krasheninnikov AV. Structural defects in graphene. ACS Nano 2010;5:26–41.
[375] Jamgotchian H, Colignon Y, Ealet B, Parditka B, Hoarau J-Y, Girardeaux C, et al. Silicene on Ag(1 1 1): domains and local defects of the observed
superstructures. J Phys: Conf Ser 2014;491:012001.
[376] Liu H, Feng H, Du Y, Chen J, Wu K, Zhao J. Point defects in epitaxial silicene on Ag(1 1 1) surface; 2015. Available from: ArXiv:151208860.
[377] Gao J, Zhang J, Liu H, Zhang Q, Zhao J. Structures, mobilities, electronic and magnetic properties of point defects in silicene. Nanoscale
2013;5:9785–92.
[378] Sahin H, Sivek J, Li S, Partoens B, Peeters FM. Stone–Wales defects in silicene: formation, stability, and reactivity of defect sites. Phys Rev B
2013;88:045434.
[379] Manjanath A, Singh AK. Low formation energy and kinetic barrier of Stone–Wales defect in infinite and finite silicene. Chem Phys Lett 2014;592:52–5.
[380] Haldar S, Amorim RG, Sanyal B, Scheicher RH, Rocha AR. Energetic stability, STM fingerprints and electronic transport properties of defects in
graphene and silicene. RSC Adv 2016;6:6702–8.
[381] Dong H, Fang D, Gong B, Zhang Y, Zhang E, Zhang S. Electronic and magnetic properties of zigzag silicene nanoribbons with Stone–Wales defects. J
Appl Phys 2015;117:064307.
[382] Iordanidou K, Houssa M, van den Broek B, Pourtois G, Afanas’ev VV, Stesmans A. Impact of point defects on the electronic and transport properties of
silicene nanoribbons. J Phys: Condens Matter 2016;28:035302.
[383] Li R, Han Y, Hu T, Dong J, Kawazoe Y. Self-healing monovacancy in low-buckled silicene studied by first-principles calculations. Phys Rev B
2014;90:045425.
[384] Li S, Wu Y, Tu Y, Wang Y, Jiang T, Liu W, et al. Defects in silicene: vacancy clusters, extended line defects, and di-adatoms. Sci Rep 2015;5:7881.
[385] Özçelik VO, Gurel HH, Ciraci S. Self-healing of vacancy defects in single-layer graphene and silicene. Phys Rev B 2013;88:045440.
[386] Song Y-L, Zhang Y, Zhang J-M, Lu D-B, Xu K-W. First-principles study of the structural and electronic properties of armchair silicene nanoribbons with
vacancies. J Mol Struct 2011;990:75–8.
[387] Li H-p, Zhang R-q. Vacancy-defect-induced diminution of thermal conductivity in silicene. Europhys Lett 2012;99:36001.
[388] Gürel HH, Özçelik VO, Ciraci S. Dissociative adsorption of molecules on graphene and silicene. J Phys Chem C 2014;118:27574–82.
[389] Hu W, Wu X, Li Z, Yang J. Porous silicene as a hydrogen purification membrane. Phys Chem Chem Phys 2013;15:5753–7.
[390] Hu W, Wu X, Li Z, Yang J. Helium separation via porous silicene based ultimate membrane. Nanoscale 2013;5:9062–6.
[391] Li R, Han Y, Dong J. Substrate effects on the monovacancies of silicene: studied from first principle methods. Phys Chem Chem Phys
2015;17:22969–76.
[392] Li Z, Feng H, Zhuang J, Pu N, Wang L, Xu X, et al. Metal–silicene interaction studied by scanning tunneling microscopy. J Phys: Condens Matter
2016;28:034002.
[393] Pedersen TG, Flindt C, Pedersen J, Mortensen NA, Jauho A-P, Pedersen K. Graphene antidot lattices: designed defects and spin qubits. Phys Rev Lett
2008;100:136804.
[394] Dvorak M, Oswald W, Wu Z. Bandgap opening by patterning graphene. Sci Rep 2013;3:2289.
[395] Ye X-S, Shao Z-G, Zhao H, Yang L, Wang C-L. Electronic and optical properties of silicene nanomeshes. RSC Adv 2014;4:37998–8003.
[396] Gong L, Xiu SL, Zheng MM, Zhao P, Zhang Z, Liang YY, et al. Electronic properties of silicene superlattices: roles of degenerate perturbation and
inversion symmetry breaking. J Mater Chem C 2014;2:8773–9.
[397] Pan F, Wang Y, Jiang K, Ni Z, Ma J, Zheng J, et al. Silicene nanomesh. Sci Rep 2015;5:9075.
[398] Mehdi Aghaei S, Calizo I. Band gap tuning of armchair silicene nanoribbons using periodic hexagonal holes. J Appl Phys 2015;118:104304.
[399] Jia T-T, Zheng M-M, Fan X-Y, Su Y, Li S-J, Liu H-Y, et al. Band gap on/off switching of silicene superlattice. J Phys Chem C 2015;119:20747–54.
[400] Lima MP, Fazzio A, da Silva AJR. Interfaces between buckling phases in silicene: ab initio density functional theory calculations. Phys Rev B
2013;88:235413.
[401] Le NB, Huan TD, Woods LM. Tunable spin-dependent properties of zigzag silicene nanoribbons. Phys Rev Appl 2014;1:054002.
[402] Yang M, Chen D-H, Wang R-Q, Bai Y-K. Spin helical states and spin transport of the line defect in silicene lattice. Phys Lett A 2015;379:396–400.
[403] Miwa RH, Kagimura R, Lima MP, Fazzio A. Valley Hall effect in silicene and hydrogenated silicene ruled by grain boundaries: an ab initio investigation.
Phys Rev B 2015;91:205442.
[404] Ghosh D, Parida P, Pati SK. Stable line defects in silicene. Phys Rev B 2015;92:195136.
[405] Zhang J, Zhao J. Structures and electronic properties of symmetric and nonsymmetric graphene grain boundaries. Carbon 2013;55:151–9.
[406] Özçelik VO, Kecik D, Durgun E, Ciraci S. Adsorption of group IV elements on graphene, silicene, germanene, and stanene: Dumbbell formation. J Phys
Chem C 2015;119:845–53.
[407] Kaloni TP, Cheng YC, Schwingenschlögl U. Hole doped Dirac states in silicene by biaxial tensile strain. J Appl Phys 2013;113:104305.
[408] Wang Y, Ding Y. Strain-induced self-doping in silicene and germanene from first-principles. Solid State Commun 2013;155:6–11.
[409] Lin X, Ni J. Dirac points and van Hove singularities of silicene under uniaxial strain. J Appl Phys 2015;117:164305.
[410] Yan J-A, Gao S-P, Stein R, Coard G. Tuning the electronic structure of silicene and germanene by biaxial strain and electric field. Phys Rev B
2015;91:245403.
[411] Mohan B, Kumar A, Ahluwalia PK. Electronic and optical properties of silicene under uni-axial and bi-axial mechanical strains: a first principle study.
Physica E 2014;61:40–7.
[412] Qin R, Wang C-H, Zhu W, Zhang Y. First-principles calculations of mechanical and electronic properties of silicene under strain. AIP Adv
2012;2:022159.
[413] Qin R, Zhu W, Zhang Y, Deng X. Uniaxial strain-induced mechanical and electronic property modulation of silicene. Nanoscale Res Lett 2014;9:521.
[414] Lew Yan Voon LC, Lopez-Bezanilla A, Wang J, Zhang Y, Willatxen M. Effective Hamiltonians for phosphorene and silicene. New J Phys
2015;17:025004.
[415] Restrepo OD, Mishra R, Goldberger JE, Windl W. Tunable gaps and enhanced mobilities in strain-engineered silicane. J Appl Phys 2014;115:033711.
[416] Shu H, Wang S, Li Y, Yip J, Wang J. Tunable electronic and optical properties of monolayer silicane under tensile strain: a many-body study. J Chem
Phys 2014;141:064707.
[417] Pei Q-X, Zhang Y-W, Sha Z-D, Shenoy VB. Tuning the thermal conductivity of silicene with tensile strain and isotopic doping: a molecular dynamics
study. J Appl Phys 2013;114:033526.
[418] Hu M, Zhang X, Poulikakos D. Anomalous thermal response of silicene to uniaxial stretching. Phys Rev B 2013;87:195417.
[419] Xie H, Ouyang T, Germaneau É, Qin G, Hu M, Bao H. Large tunability of lattice thermal conductivity of monolayer silicene via mechanical strain. Phys
Rev B 2016;93:075404.
[420] Lin Y-M, Dimitrakopoulos C, Jenkins KA, Farmer DB, Chiu H-Y, Grill A, et al. 100-GHz transistors from wafer-scale epitaxial graphene. Science
2010;327:662.
[421] Moon JS, Curtis D, Hu M, Wong D, McGuire C, Campbell PM, et al. Epitaxial-graphene RF field-effect transistors on Si-face 6H-SiC substrates. IEEE
Electron Device Lett 2009;30:650–2.
[422] Kedzierski J, Pei-Lan H, Healey P, Wyatt PW, Keast CL, Sprinkle M, et al. Epitaxial graphene transistors on SiC substrates. IEEE Trans Electron Devices
2008;55:2078–85.
J. Zhao et al. / Progress in Materials Science 83 (2016) 24–151 149

[423] Kim KK, Hsu A, Jia X, Kim SM, Shi Y, Hofmann M, et al. Synthesis of monolayer hexagonal boron nitride on Cu foil using chemical vapor deposition.
Nano Lett 2011;12:161–6.
[424] Decker R, Wang Y, Brar VW, Regan W, Tsai HZ, Wu Q, et al. Local electronic properties of graphene on a BN substrate via scanning tunneling
microscopy. Nano Lett 2011;11:2291–5.
[425] Xue J, Sanchez-Yamagishi J, Bulmash D, Jacquod P, Deshpande A, Watanabe K, et al. Scanning tunnelling microscopy and spectroscopy of ultra-flat
graphene on hexagonal boron nitride. Nat Mater 2011;10:282–5.
[426] Gannett W, Regan W, Watanabe K, Taniguchi T, Crommie MF, Zettl A. Boron nitride substrates for high mobility chemical vapor deposited graphene.
Appl Phys Lett 2011;98:242105.
[427] Dean CR, Young AF, Meric I, Lee C, Wang L, Sorgenfrei S, et al. Boron nitride substrates for high-quality graphene electronics. Nat Nanotechnol
2010;5:722–6.
[428] Liu H, Gao J, Zhao J. Silicene on substrates: a way to preserve or tune its electronic properties. J Phys Chem C 2013;117:10353–9.
[429] Li L, Wang X, Zhao X, Zhao M. Moiré superstructures of silicene on hexagonal boron nitride: a first-principles study. Phys Lett A 2013;377:2628–32.
[430] Gao N, Li JC, Jiang Q. Bandgap opening in silicene: effect of substrates. Chem Phys Lett 2014;592:222–6.
[431] Scalise E, Houssa M, Cinquanta E, Grazianetti C, van den Broek B, Pourtois G, et al. Engineering the electronic properties of silicene by tuning the
composition of MoX2 and GaX (X = S, Se, Te) chalcogenide templates. 2D Mater 2014;1:011010.
[432] Li L, Zhao M. 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;118:19129–38.
[433] Jiajie Z, Schwingenschlögl U. Silicene on MoS2: role of the van der Waals interaction. 2D Mater 2015;2:045004.
[434] Gao N, Li JC, Jiang Q. Tunable band gaps in silicene–MoS2 heterobilayers. Phys Chem Chem Phys 2014;16:11673–8.
[435] Sattar S, Hoffmann R, Schwingenschlögl U. Solid argon as a possible substrate for quasi-freestanding silicene. New J Phys 2014;16:065001.
[436] Zhu J, Schwingenschlogl U. Structural and electronic properties of silicene on MgX2 (X = Cl, Br, and I). ACS Appl Mater Interfaces 2014;6:11675–81.
[437] Li S, Wu Y, Liu W, Zhao Y. Control of band structure of van der Waals heterostructures: silicene on ultrathin silicon nanosheets. Chem Phys Lett
2014;609:161–6.
[438] Kokott S, Matthes L, Bechstedt F. Silicene on hydrogen-passivated Si(1 1 1) and Ge(1 1 1) substrates. Phys Status Solidi RRL 2013;7:538–41.
[439] Ding Y, Wang Y. Electronic structures of silicene/GaS heterosheets. Appl Phys Lett 2013;103:043114.
[440] Li Y, Chen Z. XH/p (X = C, Si) Interactions in graphene and silicene: weak in strength, strong in tuning band structures. J Phys Chem Lett
2013;4:269–75.
[441] Zhang R-w, Zhang C-w, Ji W-x, Hu S-j, Yan S-s, Li S-s, et al. Silicane as an inert substrate of silicene: a promising candidate for FET. J Phys Chem C
2014;118:25278–83.
[442] Filippone F. Interaction of silicene with beta-Si3N4(0 0 0 1)/Si(1 1 1) substrate; energetics and electronic properties. J Phys: Condens Matter
2014;26:395009.
[443] Houssa M, Scalise E, van den Broek B, Lu A, Pourtois G, Afanas’ev VV, et al. Interaction of silicene and germanene with non-metallic substrates. J Phys:
Conf Ser 2015;574:012015.
[444] Zhu J, Schwingenschlogl U. Stability and electronic properties of silicene on WSe2. J Mater Chem C 2015;3:3946–53.
p
[445] Huang Z-Q, Chou B-H, Hsu C-H, Chuang F-C, Lin H, Bansil A. Tunable topological electronic structure of silicene on a semiconducting Bi/Si(1 1 1)-
p
3  3 substrate. Phys Rev B 2014;90:245433.
[446] Houssa M, van den Broek B, Scalise E, Pourtois G, Afanas’ev VV, Stesmans A. An electric field tunable energy band gap at silicene/(0 0 0 1) ZnS
interfaces. Phys Chem Chem Phys 2013;15:3702–5.
[447] Oostinga JB, Heersche HB, Liu X, Morpurgo AF, Vandersypen LM. Gate-induced insulating state in bilayer graphene devices. Nat Mater 2008;7:151–7.
[448] Chen K, Wan X, Xu J-B. Controllable modulation of the electronic properties of graphene and silicene by interface engineering and pressure. J Mater
Chem C 2013;1:4869.
[449] Zhang J, Hong Y, Tong Z, Xiao Z, Bao H, Yue Y. Molecular dynamics study of interfacial thermal transport between silicene and substrates. Phys Chem
Chem Phys 2015;17:23704–10.
[450] Liu Z, Song L, Zhao S, Huang J, Ma L, Zhang J, et al. Direct growth of graphene/hexagonal boron nitride stacked layers. Nano Lett 2011;11:2032–7.
[451] Kim SM, Hsu A, Araujo PT, Lee Y-H, Palacios T, Dresselhaus M, et al. Synthesis of patched or stacked graphene and h-BN flakes: a route to hybrid
structure discovery. Nano Lett 2013;13:933–41.
[452] Britnell L, Gorbachev RV, Jalil R, Belle BD, Schedin F, Mishchenko A, et al. Field-effect tunneling transistor based on vertical graphene heterostructures.
Science 2012;335:947–50.
[453] Shi Y, Zhou W, Lu A-Y, Fang W, Lee Y-H, Hsu AL, et al. Van der Waals epitaxy of MoS2 layers using graphene as growth templates. Nano Lett
2012;12:2784–91.
[454] Britnell L, Gorbachev RV, Jalil R, Belle BD, Schedin F, Katsnelson MI, et al. Electron tunneling through ultrathin boron nitride crystalline barriers. Nano
Lett 2012;12:1707–10.
[455] Bertolazzi S, Krasnozhon D, Kis A. Nonvolatile memory cells based on MoS2/graphene heterostructures. ACS Nano 2013;7:3246–52.
[456] Yu WJ, Li Z, Zhou H, Chen Y, Wang Y, Huang Y, et al. Vertically stacked multi-heterostructures of layered materials for logic transistors and
complementary inverters. Nat Mater 2013;12:246–52.
[457] Georgiou T, Jalil R, Belle BD, Britnell L, Gorbachev RV, Morozov SV, et al. Vertical field-effect transistor based on graphene-WS2 heterostructures for
flexible and transparent electronics. Nat Nanotechnol 2013;8:100–3.
[458] Hu W, Li Z, Yang J. Structural, electronic, and optical properties of hybrid silicene and graphene nanocomposite. J Chem Phys 2013;139:154704.
[459] Cai Y, Chuu C-P, Wei CM, Chou MY. Stability and electronic properties of two-dimensional silicene and germanene on graphene. Phys Rev B
2013;88:245408.
[460] Wang Y, Ni Z, Liu Q, Quhe R, Zheng J, Ye M, et al. All-metallic vertical transistors based on stacked dirac materials. Adv Funct Mater 2015;25:68–77.
[461] Berdiyorov GR, Bahlouli H, Peeters FM. Theoretical study of electronic transport properties of a graphene–silicene bilayer. J Appl Phys
2015;117:225101.
[462] Peymanirad F, Neek-Amal M, Beheshtian J, Peeters FM. Graphene–silicene bilayer: a nanocapacitor with permanent dipole and piezoelectricity effect.
Phys Rev B 2015;92:155113.
[463] Liu B, Baimova JA, Reddy CD, Law AW, Dmitriev SV, Wu H, et al. Interfacial thermal conductance of a silicene/graphene bilayer heterostructure and
the effect of hydrogenation. ACS Appl Mater Interfaces 2014;6:18180–8.
[464] Berdiyorov GR, Neek-Amal M, Peeters FM, van Duin ACT. Stabilized silicene within bilayer graphene: a proposal based on molecular dynamics and
density-functional tight-binding calculations. Phys Rev B 2014;89:024107.
[465] Yu S, Li XD, Wu SQ, Wen YH, Zhou S, Zhu ZZ. Novel electronic structures of superlattice composed of graphene and silicene. Mater Res Bull
2014;50:268–72.
[466] Kamal C, Chakrabarti A, Banerjee A. Ab initio investigation on hybrid graphite-like structure made up of silicene and boron nitride. Phys Lett A
2014;378:1162–9.
[467] Kaloni TP, Tahir M, Schwingenschlogl U. Quasi free-standing silicene in a superlattice with hexagonal boron nitride. Sci Rep 2013;3:3192.
[468] Li S-s, Zhang C-w, Ji W-x. Novel electronic properties in silicene on MoSe2 monolayer: an excellent prediction for FET. Mater Chem Phys
2015;164:150–6.
[469] Li SS, Zhang CW, Yan SS, Hu SJ, Ji WX, Wang PJ, et al. Novel band structures in silicene on monolayer zinc sulfide substrate. J Phys: Condens Matter
2014;26:395003.
150 J. Zhao et al. / Progress in Materials Science 83 (2016) 24–151

[470] Zhang R-w, Zhang C-w, Ji W-x, Ren M-j, Li F, Yuan M. First-principles prediction on silicene-based heterobilayers as a promising candidate for FET.
Mater Chem Phys 2015;156:89–94.
[471] Li X, Wu S, Zhou S, Zhu Z. Structural and electronic properties of germanene/MoS2 monolayer and silicene/MoS2 monolayer superlattices. Nanoscale
Res Lett 2014;9:110.
[472] Kou L, Ma Y, Yan B, Tan X, Chen C, Smith SC. Encapsulated silicene: a robust large-gap topological insulator. ACS Appl Mater Interfaces
2015;7:19226–33.
[473] Fang D-Q, Zhang S-L, Xu H. Tuning the electronic and magnetic properties of zigzag silicene nanoribbons by edge hydrogenation and doping. RSC Adv
2013;3:24075.
[474] Luan H-X, Zhang C-W, Zheng F-B, Wang P-J. First-principles study of the electronic properties of B/N atom doped silicene nanoribbons. J Phys Chem C
2013;117:13620–6.
[475] Ma L, Zhang J-M, Xu K-W, Ji V. Structural and electronic properties of substitutionally doped armchair silicene nanoribbons. Physica B
2013;425:66–71.
[476] Zheng F-b, Zhang C-w, Yan S-s, Li F. Novel electronic and magnetic properties in N or B doped silicene nanoribbons. J Mater Chem C 2013;1:2735.
[477] Dong Y-J, Wang X-F, Vasilopoulos P, Zhai M-X, Wu X-M. Half-metallicity in aluminum-doped zigzag silicene nanoribbons. J Phys D Appl Phys
2014;47:105304.
[478] Ma L, Zhang J-M, Xu K-W, Ji V. Nitrogen and boron substitutional doped zigzag silicene nanoribbons: ab initio investigation. Physica E 2014;60:112–7.
[479] Chen X, Ni J. Predicted ferromagnetism in hole doped armchair nanoribbons: a first principles study. Chem Phys Lett 2013;555:173–7.
[480] Chen J, Wang XF, Vasilopoulos P, Chen AB, Wu JC. Single and multiple doping effects on charge transport in zigzag silicene nanoribbons.
Chemphyschem: Eur J Chem Phys Phys Chem 2014;15:2701–6.
[481] Zhang J-M, Song W-T, Xu K-W, Ji V. The study of the P doped silicene nanoribbons with first-principles. Comput Mater Sci 2014;95:429–34.
[482] Lopez-Bezanilla A. Substitutional doping widens silicene gap. J Phys Chem C 2014;118:18788–92.
[483] Song YL, Zhang Y, Zhang JM, Lu DB, Xu KW. Modulation of the electronic and magnetic properties of the silicene nanoribbons by a single C chain. Eur
Phys J B 2010;79:197–202.
[484] Song Y-L, Zhang J-M, Lu D-B, Xu K-W. Structural and electronic properties of a single C chain doped zigzag silicene nanoribbon. Physica E
2013;53:173–7.
[485] Song Y-L, Zhang J-M, Lu D-B, Xu K-W. Structural and electromagnetic properties of double C chains decorated zigzag silicene nanoribbon. Physica E
2014;56:205–10.
[486] Das R, Chowdhury S, Majumdar A, Jana D. Optical properties of P and Al doped silicene: a first principles study. RSC Adv 2015;5:41–50.
[487] Guo Y, Zhou S, Bai Y, Zhao J. Tunable thermal conductivity of silicene by germanium doping. J Supercond Nov Magn 2015;29:717–20.
[488] Tan X, Li F, Chen Z. Metallic BSi3 silicene and its one-dimensional derivatives: unusual nanomaterials with planar aromatic D6h six-membered silicon
rings. J Phys Chem C 2014;118:25825–35.
[489] Ding Y, Wang Y. Density functional theory study of the silicene-like SiX and XSi3(X = B, C, N, Al, P) honeycomb lattices: the various buckled structures
and versatile electronic properties. J Phys Chem C 2013;117:18266–78.
[490] Drissi LB, Saidi EH, Bousmina M, Fassi-Fehri O. DFT investigations of the hydrogenation effect on silicene/graphene hybrids. J Phys: Condens Matter
2012;24:485502.
[491] Lin X, Lin S, Xu Y, Hakro AA, Hasan T, Zhang B, et al. Ab initio study of electronic and optical behavior of two-dimensional silicon carbide. J Mater
Chem C 2013;1:2131.
[492] Liu G, Wu MS, Ouyang CY, Xu B. Structural and electronic evolution from SiC sheet to silicene. Int J Mod Phys B 2013;27:1350188.
[493] Drissi LB, Ramadan FZ. Many body effects study of electronic & optical properties of silicene–graphene hybrid. Physica E 2015;68:38–41.
[494] Qin X, Liu Y, Li X, Xu J, Chi B, Zhai D, et al. Origin of Dirac cones in SiC Silagraphene: a combined density functional and tight-binding study. J Phys
Chem Lett 2015;6:1333–9.
[495] Padilha JE, Seixas L, Pontes RB, da Silva AJR, Fazzio A. Quantum spin Hall effect in a disordered hexagonal SixGe1x alloy. Phys Rev B 2013;88:201106.
[496] Zhou H, Zhao M, Zhang X, Dong W, Wang X, Bu H, et al. First-principles prediction of a new Dirac-fermion material: silicon germanide monolayer. J
Phys: Condens Matter 2013;25:395501.
[497] Zhang R-w, Zhang C-w, Li S-s, Ji W-x, Wang P-j, Li F, et al. Design of half-metallic ferromagnetism in germanene/silicene hybrid sheet. Solid State
Commun 2014;191:49–53.
[498] Ezawa M. A topological insulator and helical zero mode in silicene under an inhomogeneous electric field. New J Phys 2012;14:033003.
[499] Liu W, Wang ZF, Shi QW, Yang J, Liu F. Band-gap scaling of graphene nanohole superlattices. Phys Rev B 2009;80:233405.
[500] Yu B, Chang L, Ahmed S, Wang H, Bell S, Yang C-Y, et al. FinFET scaling to l0 nm gate length. In: IEEE int electron devices meeting. San Francisco, CA,
USA. p. 251.
[501] Doris B, Ieong M, Zhu T, Zhang Y, Steen M, Natzle W, et al. Device design considerations for ultra-thin SOI MOSFETs. In: IEEE int electron devices
meeting. Washington, DC, USA: IEEE; 2003. p. 27.3.1-.3.4.
[502] Franklin AD, Luisier M, Han S-J, Tulevski G, Breslin CM, Gignac L, et al. Sub-10 nm carbon nanotube transistor. Nano Lett 2012;12:758.
[503] Li H, Wang L, Liu Q, Zheng J, Mei W-N, Gao Z, et al. High performance silicene nanoribbon field effect transistors with current saturation. Eur Phys J B
2012;85:274.
[504] Lin Y-M, Jenkins KA, Valdes-Garcia A, Small JP, Farmer DB, Avouris P. Operation of graphene transistors at gigahertz frequencies. Nano Lett
2009;9:422–6.
[505] Yan Q, Huang B, Yu J, Zheng F, Zang J, Wu J, et al. Intrinsic current–voltage characteristics of graphene nanoribbon transistors and effect of edge
doping. Nano Lett 2007;7:1469–73.
[506] Franklin AD, Chen Z. Length scaling of carbon nanotube transistors. Nat Nanotechnol 2010;5:858–62.
[507] Colinge J-P, Lee C-W, Afzalian A, Akhavan ND, Yan R, Ferain I, et al. Nanowire transistors without junctions. Nat Nanotechnol 2010;5:225–9.
[508] Sacconi F, Persson MP, Povolotskyi M, Latessa L, Pecchia A, Gagliardi A, et al. Electronic and transport properties of silicon nanowires. J Comput
Electron 2007;6:329–33.
[509] Yi KS, Trivedi K, Floresca HC, Yuk H, Hu W, Kim MJ. Room-temperature quantum confinement effects in transport properties of ultrathin Si nanowire
field-effect transistors. Nano Lett 2011;11:5465–70.
[510] Yang XF, Liu YS, Feng JF, Wang XF, Zhang CW, Chi F. Transport properties of bare and hydrogenated zigzag silicene nanoribbons: negative differential
resistances and perfect spin-filtering effects. J Appl Phys 2014;116:124312.
[511] Son Y-W, Cohen ML, Louie SG. Half-metallic graphene nanoribbons. Nature 2006;444:347–9.
[512] Kang J, Wu F, Li J. Doping induced spin filtering effect in zigzag graphene nanoribbons with asymmetric edge hydrogenation. Appl Phys Lett
2011;98:083109.
[513] Deng X, Zhang Z, Tang G, Fan Z, Zhu H, Yang C. Edge contact dependent spin transport for n-type doping zigzag-graphene with asymmetric edge
hydrogenation. Sci Rep 2014;4:4038.
[514] Xu C, Luo G, Liu Q, Zheng J, Zhang Z, Nagase S, et al. Giant magnetoresistance in silicene nanoribbons. Nanoscale 2012;4:3111–7.
[515] Muñoz-Rojas F, Fernández-Rossier J, Palacios JJ. Giant magnetoresistance in ultrasmall graphene based devices. Phys Rev Lett 2009;102:136810.
[516] Lado JL, Fernández-Rossier J. Magnetic edge anisotropy in graphene like honeycomb crystals. Phys Rev Lett 2014;113:027203.
[517] Pan F, Quhe R, Ge Q, Zheng J, Ni Z, Wang Y, et al. Gate-induced half-metallicity in semihydrogenated silicene. Physica E 2014;56:43–7.
[518] Yokoyama T. Controllable valley and spin transport in ferromagnetic silicene junctions. Phys Rev B 2013;87:241409.
[519] Yokoyama T. Spin and valley transports in junctions of Dirac fermions. New J Phys 2014;16:085005.
[520] Xiao D, Yao W, Niu Q. Valley-contrasting physics in graphene: magnetic moment and topological transport. Phys Rev Lett 2007;99:236809.
J. Zhao et al. / Progress in Materials Science 83 (2016) 24–151 151

[521] Ouyang Y, Guo J. A theoretical study on thermoelectric properties of graphene nanoribbons. Appl Phys Lett 2009;94:263107.
[522] Huang W, Da H, Liang G. Thermoelectric performance of MX2 (M = Mo, W; X = S, Se) monolayers. J Appl Phys 2013;113:104304.
[523] Wu J, Schmidt H, Amara KK, Xu X, Eda G, Özyilmaz B. Large thermoelectricity via variable range hopping in chemical vapor deposition grown single-
layer MoS2. Nano Lett 2014;14:2730–4.
[524] Yang K, Cahangirov S, Cantarero A, Rubio A, D’Agosta R. Thermoelectric properties of atomically thin silicene and germanene nanostructures. Phys Rev
B 2014;89:125403.
[525] Zberecki K, Wierzbicki M, Barnaś J, Swirkowicz R. Thermoelectric effects in silicene nanoribbons. Phys Rev B 2013;88:115404.
[526] Pan L, Liu HJ, Tan XJ, Lv HY, Shi J, Tang XF, et al. Thermoelectric properties of armchair and zigzag silicene nanoribbons. Phys Chem Chem Phys
2012;14:13588–93.
[527] Sadeghi H, Sangtarash S, Lambert CJ. Enhanced thermoelectric efficiency of porous silicene nanoribbons. Sci Rep 2015;5:9514.
[528] Zhao W, Guo ZX, Zhang Y, Ding JW, Zheng XJ. Enhanced thermoelectric performance of defected silicene nanoribbons. Solid State Commun
2016;227:1–8.
[529] Wierzbicki M, Barnas J, Swirkowicz R. Zigzag nanoribbons of two-dimensional silicene-like crystals: magnetic, topological and thermoelectric
properties. J Phys: Condens Matter 2015;27:485301.
[530] Yan Y, Wu H, Jiang F, Zhao H. Enhanced thermopower of gated silicene. Eur Phys J B 2013;86:457.
[531] Zberecki K, Swirkowicz R, Barnaś J. Spin effects in thermoelectric properties of Al- and P-doped zigzag silicene nanoribbons. Phys Rev B
2014;89:165419.
[532] Zberecki K, Swirkowicz R, Wierzbicki M, Barnas J. Enhanced thermoelectric efficiency in ferromagnetic silicene nanoribbons terminated with
hydrogen atoms. Phys Chem Chem Phys 2014;16:12900–8.
[533] An R-L, Wang X-F, Vasilopoulos P, Liu Y-S, Chen A-B, Dong Y-J, et al. Vacancy effects on electric and thermoelectric properties of zigzag silicene
nanoribbons. J Phys Chem C 2014;118:21339–46.
[534] Zberecki K, Swirkowicz R, Barnaś J. Thermoelectric properties of zigzag silicene nanoribbons doped with Co impurity atoms. J Magn Magn Mater
2015;393:305–9.
[535] Xu L, Wang X-F, Zhou L, Yang Z-Y. Adsorption of Ti atoms on zigzag silicene nanoribbons: influence on electric, magnetic, and thermoelectric
properties. J Phys D Appl Phys 2015;48:215306.
[536] Ping Niu Z, Dong S. Valley and spin thermoelectric transport in ferromagnetic silicene junctions. Appl Phys Lett 2014;104:202401.
[537] Fu H-H, Wu D-D, Zhang Z-Q, Gu L. Spin-dependent Seebeck effect, thermal colossal magnetoresistance and negative differential thermoelectric
resistance in zigzag silicene nanoribbon heterojunction. Sci Rep 2015;5:10547.
[538] Wierzbicki M, Barnaś J, Swirkowicz R. Thermoelectric properties of silicene in the topological- and band-insulator states. Phys Rev B 2015;91:165417.
[539] Feng J-w, Liu Y-j, Wang H-x, Zhao J-x, Cai Q-h, Wang X-z. Gas adsorption on silicene: a theoretical study. Comput Mater Sci 2014;87:218–26.
[540] Chandiramouli R, Srivastava A, Nagarajan V. NO adsorption studies on silicene nanosheet: DFT investigation. Appl Surf Sci 2015;351:662–72.
[541] Hu W, Xia N, Wu X, Li Z, Yang J. Silicene as a highly sensitive molecule sensor for NH3, NO and NO2. Phys Chem Chem Phys 2014;16:6957–62.
[542] Osborn TH, Farajian AA. Silicene nanoribbons as carbon monoxide nanosensors with molecular resolution. Nano Res 2014;7:945–52.
[543] Amorim RG, Scheicher RH. Silicene as a new potential DNA sequencing device. Nanotechnology 2015;26:154002.
[544] Sadeghi H, Bailey S, Lambert CJ. Silicene-based DNA nucleobase sensing. Appl Phys Lett 2014;104:103104.
[545] Jose D, Datta A. Structures and electronic properties of silicene clusters: a promising material for FET and hydrogen storage. Phys Chem Chem Phys
2011;13:7304–11.
[546] Wang Y, Zheng R, Gao H, Zhang J, Xu B, Sun Q, et al. Metal adatoms-decorated silicene as hydrogen storage media. Int J Hydrogen Energy
2014;39:14027–32.
[547] Lochan RC, Head-Gordon M. Computational studies of molecular hydrogen binding affinities: the role of dispersion forces, electrostatics, and orbital
interactions. Phys Chem Chem Phys 2006;8:1357–70.
[548] Hussain T, Kaewmaraya T, Chakraborty S, Ahuja R. Functionalization of hydrogenated silicene with alkali and alkaline earth metals for efficient
hydrogen storage. Phys Chem Chem Phys 2013;15:18900–5.
[549] Hussain T, Chakraborty S, De Sarkar A, Johansson B, Ahuja R. Enhancement of energy storage capacity of Mg functionalized silicene and silicane under
external strain. Appl Phys Lett 2014;105:123903.
[550] Hussain T, Chakraborty S, Ahuja R. Metal-functionalized silicene for efficient hydrogen storage. ChemPhysChem 2013;14:3463–6.
[551] Li F, Zhang C-w, Luan H-x, Wang P-j. First-principles study of hydrogen storage on Li-decorated silicene. J Nanopart Res 2013;15:1972.
[552] Wang J, Li J, Li S-S, Liu Y. Hydrogen storage by metalized silicene and silicane. J Appl Phys 2013;114:124309.
[553] Song EH, Yoo SH, Kim JJ, Lai SW, Jiang Q, Cho SO. External electric field induced hydrogen storage/release on calcium-decorated single-layer and
bilayer silicene. Phys Chem Chem Phys 2014;16:23985–92.
[554] Tritsaris GA, Kaxiras E, Meng S, Wang E. Adsorption and diffusion of lithium on layered silicon for Li-ion storage. Nano Lett 2013;13:2258–63.
[555] Deng J, Liu JZ, Medhekar NV. Enhanced lithium adsorption and diffusion on silicene nanoribbons. RSC Adv 2013;3:20338.
[556] Tan X, Cabrera CR, Chen Z. Metallic BSi3 silicene: a promising high capacity anode material for lithium-ion batteries. J Phys Chem C
2014;118:25836–43.
[557] Kulish VV, Malyi OI, Ng MF, Chen Z, Manzhos S, Wu P. Controlling Na diffusion by rational design of Si-based layered architectures. Phys Chem Chem
Phys 2014;16:4260–7.
[558] Hwang Y, Yun K-H, Chung Y-C. Carbon-free and two-dimensional cathode structure based on silicene for lithium–oxygen batteries: a first-principles
calculation. J Power Sources 2015;275:32–7.
[559] Chen X, Wu Y, Wu Z, Han Y, Xu S, Wang L, et al. High-quality sandwiched black phosphorus heterostructure and its quantum oscillations. Nat
Commun 2015;6:7315.
[560] Xu Y, Yan B, Zhang H-J, Wang J, Xu G, Tang P, et al. Large-gap quantum spin Hall insulators in tin films. Phys Rev Lett 2013;111:136804.
[561] Wu S-C, Shan G, Yan B. Prediction of near-room-temperature quantum anomalous Hall effect on honeycomb materials. Phys Rev Lett
2014;113:256401.
[562] Dávila ME, Xian L, Cahangirov S, Rubio A, Le Lay G. Germanene: a novel two-dimensional germanium allotrope akin to graphene and silicene. New J
Phys 2014;16:095002.
[563] Davila ME, Le Lay G. Few layer epitaxial germanene: a novel two-dimensional Dirac material. Sci Rep 2016;6:20714.
[564] Li L, S.-z. Lu, Pan J, Qin Z, Wang Y-q, Wang Y, et al. Buckled germanene formation on Pt(1 1 1). Adv Mater 2014;26:4820–4.
[565] d’Acapito F, Torrengo S, Xenogiannopoulou E, Tsipas P, Marquez Velasco J, Tsoutsou D, et al. Evidence for germanene growth on epitaxial hexagonal
(h)-AlN on Ag(1 1 1). J Phys: Condens Matter 2016;28:045002.
[566] Zhu F-f, Chen W-j, Xu Y, Gao C-l, Guan D-d, Liu C-h, et al. Epitaxial growth of two-dimensional stanene. Nat Mater 2015;14:1020–5.
[567] Bianco E, Butler S, Jiang S, Restrepo OD, Windl W, Goldberger JE. Stability and exfoliation of germanane: a germanium graphane analogue. ACS Nano
2013;7:4414–21.
[568] Zhou X-F, Dong X, Oganov AR, Zhu Q, Tian Y, Wang H-T. Semimetallic two-dimensional boron allotrope with massless Dirac fermions. Phys Rev Lett
2014;112:085502.
[569] Li L, Wang Y, Xie S, Li XB, Wang YQ, Wu R, et al. Two-dimensional transition metal honeycomb realized: Hf on Ir(1 1 1). Nano Lett 2013;13:4671–4.

You might also like