+ All Categories
Home > Documents > Graphene-analogous low-dimensional materials

Graphene-analogous low-dimensional materials

Date post: 08-Dec-2016
Category:
Upload: zhen
View: 234 times
Download: 6 times
Share this document with a friend
173
Accepted Manuscript Graphene-Analogous Low-Dimensional Materials Qing Tang, Zhen Zhou PII: S0079-6425(13)00037-6 DOI: http://dx.doi.org/10.1016/j.pmatsci.2013.04.003 Reference: JPMS 320 To appear in: Progress in Materials Science Please cite this article as: Tang, Q., Zhou, Z., Graphene-Analogous Low-Dimensional Materials, Progress in Materials Science (2013), doi: http://dx.doi.org/10.1016/j.pmatsci.2013.04.003 This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.
Transcript
Page 1: Graphene-analogous low-dimensional materials

Accepted Manuscript

Graphene-Analogous Low-Dimensional Materials

Qing Tang, Zhen Zhou

PII: S0079-6425(13)00037-6

DOI: http://dx.doi.org/10.1016/j.pmatsci.2013.04.003

Reference: JPMS 320

To appear in: Progress in Materials Science

Please cite this article as: Tang, Q., Zhou, Z., Graphene-Analogous Low-Dimensional Materials, Progress in

Materials Science (2013), doi: http://dx.doi.org/10.1016/j.pmatsci.2013.04.003

This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers

we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and

review of the resulting proof before it is published in its final form. Please note that during the production process

errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.

Page 2: Graphene-analogous low-dimensional materials

1

Graphene-Analogous Low-Dimensional Materials

Qing Tang, Zhen Zhou*

Tianjin Key Laboratory of Metal and Molecule Based Material Chemistry, Key Laboratory of Advanced Energy

Materials Chemistry (Ministry of Education), Computational Centre for Molecular Science, Institute of New

Energy Material Chemistry, Nankai University, Tianjin 300071, P. R. China

ABSTRACT

Graphene, an atomic monolayer of carbon atoms in a honeycomb lattice realized in 2004, has

rapidly risen as the hottest star in materials science due to its exceptional properties. The explosive

studies on graphene have sparked new interests towards graphene-analogous materials. Now many

graphene-analogous materials have been fabricated from a large variety of layer and non-layer

materials. Also, many graphene-analogous materials have been designed from the computational

side. Though overshadowed by the rising graphene to some degree, graphene-analogous materials

have exceptional properties associated with low dimensionality and edge states, and bring new

breakthrough to nanomaterials science as well. In this review, we summarize the recent progress

on graphene-analogous low-dimensional materials (2D nanosheets and 1D nanoribbons) from both

experimental and computational side, and emphasis is placed on structure, properties, preparation,

and potential applications of graphene-analogous materials as well as the comparison with

graphene. The reviewed materials include strictly graphene-like planar materials (experimentally

available h-BN, silicene, and BC3 as well as computationally predicted SiC, SiC2, B, and B2C),

non-planar materials (metal dichalcogenides, metal oxides and hydroxides, graphitic-phase of ZnO,

MXene), metal coordination polymers, and organic covalent polymers. This comprehensive

review might provide a directional guide for the bright future of this emerging area.

* Corresponding author. Tel.: +86 22 23503623; fax: +86 22 23498941.

Email address: [email protected] (Z.Z.)

Page 3: Graphene-analogous low-dimensional materials

2

Contents 1. Introduction...................................................................................................................................3 2. Experimental synthesis, characterization, and theoretical methods ..............................................6

2.1 Micromechanical cleavage .................................................................................................6 2.2 Chemical exfoliation ...........................................................................................................7 2.3 Chemical vapor deposition ...............................................................................................13 2.4 Surface-assisted epitaxial growth .....................................................................................18 2.5 Synthesis of 1D nanoribbons.............................................................................................20

2.5.1 Unzipping BN nanotubes to produce BN nanoribbons ..........................................20 2.5.2 Growth of ultranarrow MoS2 and WS2 nanoribbons inside carbon nanotubes .....22

2.6 Other synthetic routes .......................................................................................................23 2.7 Characterization ...............................................................................................................25 2.8 Theoretical methods ..........................................................................................................27

3. Planar graphene analogues..........................................................................................................30 3.1 “White graphene”: BN nanosheets and nanoribbons.......................................................31

3.1.1 Comparison between BN and C .............................................................................31 3.1.2 Electronic and magnetic properties of BN nanosheets and nanoribbons ..............33 3.1.3 Band-gap modifications of BN nanosheets and nanoribbons ................................36 3.1.4 BN/graphene hybrid structures ..............................................................................46 3.1.5 Potential applications ............................................................................................55

3.2 Silicene ..............................................................................................................................59 3.2.1 Synthesis of silicene and functionalized silicene ....................................................60 3.2.2 Theoretical investigations of silicene .....................................................................62 3.2.3 Potential applications of silicene ...........................................................................69

3.3 BC3 honeycomb sheets ......................................................................................................70 4. Hypothetical planar graphene analogues ....................................................................................73

4.1 SiC Silagraphene...............................................................................................................73 4.2 SiC2 Silagraphene .............................................................................................................76 4.3 Boron sheets ......................................................................................................................78 4.4 B2C sheet ...........................................................................................................................83

5. Non-planar materials...................................................................................................................85 5.1 Metal dichalcogenides ......................................................................................................85

5.1.1 MoS2 and WS2 ........................................................................................................85 5.1.2 Other layered metal dichalcogenides.....................................................................98

5.2 Layered oxide and hydroxide nanosheets .........................................................................99 5.3 Graphitic-like phase of ZnO............................................................................................104 5.4 MXenes............................................................................................................................108

6. Two-dimensional coordination and covalent organic polymers ...............................................112 6.1 Coordination polymers ...................................................................................................112 6.2 Covalent organic polymers .............................................................................................116

7. Conclusion and prospective ......................................................................................................119 Acknowledgements.......................................................................................................................121 References.....................................................................................................................................121

Page 4: Graphene-analogous low-dimensional materials

3

1. Introduction

Graphene, a carbon honeycomb with only one-atom thickness isolated from

graphite in 2004, is the first example of true two-dimensional (2D) single-layer

atomic crystal [1]. Despite its short history, graphene shows tremendous attraction to

researchers from different fields and has risen as the most exciting star in materials

science during the past several years [ 2 ]. Its exceptional properties, such as

half-integer quantum Hall effect, ambipolar electric field effect, extremely high

carrier mobility, high thermal conductivity, high specific surface area, and the highest

strength ever mearsured, provide a fertile ground for the possible implementation of

graphene in nanodevices for a large variety of applications, and a lot of recent reviews

have been directed towards its synthesis, properties, and functionalized applications

[3-6].

As a matter of fact, the intensive studies on graphene have sparked new

discoveries towards graphene-analogous materials, comprising single layer or few

layers with compositions other than carbon [7-9]. Initially, following the same

micromechanical exfoliation methodology originally used in graphite, Novoselov et al.

[10] successfully isolated individual single layers from a large variety of layered

materials including h-BN, dichalcogenides (such as MoS2 and NbSe2) and complex

oxides (such as Ba2Sr2CaCu2Ox). Since then, many 2D materials including BC3 [11],

silicene [12], MXene (transition metal carbides and nitrides, such as Ti3C2 [13] and

Ti2C [14]), and coordination polymers (such as [Cu2Br(IN)2]n [15] and polymeric

Fe-phthalocyanine [16]), have been prepared and characterized, and new fabrication

Page 5: Graphene-analogous low-dimensional materials

4

methods, including liquid-phase exfoliation [17], electrochemical exfoliation [18],

chemical vapor deposition (CVD), and polycondensation reaction, have been

accordingly established. In addition, unique experimental skills were developed for

preparing 1D ribbons, e.g., BN nanoribbons (NRs) were produced by unzipping BN

nanotubes through plasma etching [19] or potassium intercalation [20]. Ultra-narrow

MoS2 and WS2 nanoribbons encapsulated into single- and double-walled carbon

nanotubes were synthesized via chemical reactions inside carbon nanotubes [21,22].

These novel materials have fantastic properties and promising applications. For

example, BN nanosheets are highly insulating, and exhibit superb chemical, thermal,

and oxidation stability, and possess a mechanical strength and thermal conductivity

comparable to those of graphene. These excellent properties aid the uses of layered

BN as thermal radiators, ultraviolet-light laser and emitter devices, as

thermoconductive fillers in polymer or ceramic composites, and as dielectric substrate

for graphene-based electronics [23]. Similarly, MoS2 nanosheets were explored as

high-temperature solid lubricants, nanoelectronics, electrode materials for Li-ion

batteries, and catalysts. MoS2 monolayer with a direct band gap of 1.8 eV serves as a

semiconducting rival of graphene, particularly considering that the zero-band-gap

character of the latter poses a severe drawback for its applicability in logic devices.

Actually, a MoS2-based transistor with a room-temperature mobility of more than 200

cm2/(Vs) and current on/off ratios of up to 1×108 has been fabricated, which is

demonstrated to be comparable to thin Si films or graphene nanoribbons (GNRs).

At the computational side, many graphene-like materials have been explored and

Page 6: Graphene-analogous low-dimensional materials

5

designed, and fascinating properties distinctive from graphene are predicted even in

advance of experiments. For example, in contrast to graphene, all BN monolayers

have closed-shell singlet ground states, and those with long zigzag edges have slightly

larger band gaps [24]. For BN nanoribbons, the electronic properties depend heavily

on the edge states. The ground states of both fully bare BN nanoribbons and the ones

with a bare N edge and an H-terminated B edge are half metallic. The alignment of

spin at the bare B edge is antiferromagnetic (anti-FM), while that at the bare N edge is

ferromagnetic (FM) for both the fully bare and half-bare zigzag-edged BN

nanoribbons [25]. Hydrogenation can further precisely modulate the electronic and

magnetic properties of BN nanoribbons by controlling hydrogenation ratios [26].

Zigzag SiC nanoribbons are magnetic metals, whose spin polarization originates from

the unpaired electrons localized on the ribbon edges. Interestingly, the zigzag SiC

nanoribbons narrower than ~4 nm present half-metallic behavior without the aid of

external field or chemical modification [27]. Ahead of experimental realization, Li et

al. [28] investigated the stability, magnetic and electronic properties of MoS2 sinlge

layer and nanoribbons with either zigzag or armchair-terminated edges. Zigzag

nanoribbons show ferromagnetic and metallic behavior, while armchair nanoribbons

are nonmagnetic (NM) and semiconducting. Such materials came into truth in 2010

[21]. By computations, planar tetracoordinate silicon (ptSi)-containing SiC2

silagraphene was designed [29], where each silicon atom is bonded with four carbon

atoms in a pure plane, representing the first anti-van't Hoff/Lebel species in the

Si-containing extended system. All these ptSi-containing nanomaterials are metallic.

Page 7: Graphene-analogous low-dimensional materials

6

Though overshadowed by the rising graphene to some degree,

graphene-analogous materials have exceptional properties, and bring new

breakthrough to nanomaterials science. Very recently, several good reviews have been

published on graphene-like materials, including BN, transition metal dichalcogenides,

and silicene [12,23,30]. Aside from these, many other 2D materials have also gained

recent excitement, such as Si-C binary phase, B and B-C phase, layered V2O5,

MXenes, wurtzite materials (such as ZnO), and coordination or covalent polymers. In

this review, we will summarize recent progress in all these graphene-analogous

materials from both experimental and computational side, and emphasis is placed on

structure, properties, preparation, and potential applications. It is highly expected that

this more comprehensive review might be useful for both experimental and theoretical

peers and provide a directional guide for the bright future of this emerging new area

of research.

2. Experimental synthesis, characterization, and theoretical methods

In this section we will first highlight the recent progress in general synthesis and

characterization of experimentally achieved 2D materials, and some special methods

for individual materials will be introduced in their corresponding section.

Subsequently, we will provide a brief summary on theoretical methods used in

structure modeling and property prediction of graphene-analogous materials.

2.1 Micromechanical cleavage

Micromechanical cleavage, as originally used in peeling off graphene from

graphite, can be extended to other layered materials with weak van der Waals (vdW)

Page 8: Graphene-analogous low-dimensional materials

7

forces between layers. This method requires repeatedly peeling layered materials and

followed transferring the peeled sample on top of a surface. The micromechanical

cleavage was firstly applied to isolation of h-BN, MoS2, NbSe2, and Ba2Sr2CaCu2Ox

[10]. The resulting 2D sheets are stable under ambient conditions, exhibit high crystal

quality, and are continuous on a macroscopic scale. After that, several groups have

reported synthesis of BN [31,32], MoS2, NbSe2, WSe2 [33], GaS, GaSe [34,35],

Bi2Se3 [36], and Bi2Te3 [37] nanosheets from their layered phases by using this

method, with the obtained thickness ranging from one to ten atomic layers.

Micromechanical cleavage has proven an easy and fast way of obtaining highly

crystalline atomically thin nanosheets. Nevertheless, this method also produces a large

quantity of thicker sheets, and the thinner or monolayer ones only reside in a very

minor proportion; thus this method is not scalable to mass production for potential

engineering applications. Consequently, although mechanical cleavage has shown

some success in isolating thin BN and MoS2 nanosheets, this method is practically

less explored due to its extremely low yield.

2.2 Chemical exfoliation

The critical drawback for micromechanical exfoliation is the large thickness of

obtained majority products. As an alternative method, chemically derived exfoliations,

such as liquid-phase exfoliation, and ion-intercalation induced exfoliation, have been

demonstrated to effectively isolate single layer and few layers from those thicker

structures in large quantities. These chemically derived routes, also regarded as wet

methods, have been widely adopted for synthesizing 2D materials with lateral sizes up

Page 9: Graphene-analogous low-dimensional materials

8

to several micrometers.

Liquid-phase exfoliation is regarded as a dispersion/exfoliation method, which

consists of sonicating the layered bulk materials in polar solvents, surfactant or

reaction reagents, and then exfoliating the resultant dispersions into separated thin

layers with assistance of centrifugation. The strong affinity between solvent and host

materials weakens the interlayer interactions and thus facilitates the isolation of thin

sheets upon sonication. A proper solvent should have a surface energy that matches

the energy required to overcome the vdW forces of bulk materials, and be able to

form stable dispersion with the host materials against reaggregation. The quality of

yielded materials depends heavily on solution-processing parameters such as

sonication time and centrifugation rate. Synthesis of mono- and few-layered h-BN

nanosheets from single crystalline h-BN via a chemical solution derived method was

first accomplished by Han et al. [38] in 2008. Later, Zhi et al. [39] exfoliated

large-scale BN nanosheets from BN powers dispersed in a strong polar solvent,

N,N-dimethylformamide (DMF). However, the isolated BN nanosheets suffer from

relatively low yields, corresponding to 0.01-0.03 mg/mL after 10 h sonication, which

is mainly due to the weak interactions between BN layers and the solvent molecules.

Using the liquid-phase exfoliation method, Coleman et al. [17] produced single-

and few-layered 2D nanosheets (BN, MoS2, and WS2) dispersed in various organic

solvents (Fig. 1), which were further fabricated into macroscopic films by vacuum

filtration. In their experiments, about 25-30 solvents were examined to check their

effectiveness on each material, and the promising solvents should have a surface

Page 10: Graphene-analogous low-dimensional materials

9

tension close to 40 mJ/m2, such as N-methyl-pyrrolidone (NMP) and isopropanol

(IPA). This provides useful information for exploring new solvents and solvent blends.

Under their optimized solvent conditions, the lateral sizes of the resulting nanosheets

are 50 to 1000 nm for MoS2 and WS2 and 100 to 5000 nm for BN, with the generated

concentration as high as 0.3 mg/ml for MoS2 (in NMP), 0.15 mg/ml for WS2 (in

NMP), and 0.06 mg/ml for BN (IPA). Transmission electron microscopy (TEM)

images confirmed that the exfoliated nanosheets exhibited well-preserved hexagonal

symmetry.

Fig. 1. (a) Photographs of MoS2 (in NMP), WS2 (in NMP), and BN (in IPA). (b)-(d)

High-resolution TEM (HRTEM) images of BN, MoS2, and WS2 monolayers. (e)

Photograph of BN, MoS2, and WS2 (thickness ~50 µm) films fabricated via vacuum

filtration. Reproduced from Ref. [17]. Copyright © 2011, American Association for

the Advancement of Science.

Page 11: Graphene-analogous low-dimensional materials

10

Additionally, other polar solvents, such as deionized water [ 40 ],

N-methyl-2-pyrrolidone [41], methanesulfonic acid (MSA) [42], 1,2-dichloroethane

[43], cyclohexylpyrrolidone (CHP) [44], a mixture of ethanol and water [45], sodium

cholate/water solution [46], and formamide [47,48], have been used as dispersion

media of layered materials (Table 1).

Table 1

Overview of adopted solvents and reagents for liquid-phase exfoliation of 2D

materials.

Solvents/Reagents Exfoliated materials (precursor) Ref.

N,N-dimethylformamide (DMF) BN [39]

N-methyl-2-pyrrolidone BN [41]

methanesulfonic acid (MSA) BN [42]

1,2-dichloroethane BN [43]

1,2-dichloroethane/polyphenylenevinylene BN [38]

molten hydroxides BN [49]

cyclohexylpyrrolidone (CHP) WS2, MoS2, MoSe2, MoTe2 [44]

N-methyl-pyrrolidone (NMP) MoS2, WS2, WSe2 [17,50,51]

isopropanol (IPA) BN [17]

deionized water BN [40]

N-vinyl-2-pyrrolidone MoS2 [52]

ethanol/water MoS2, WS2, BN [45]

sodium cholate solution MoS2 [46]

formamide VS2, V2O5 [47,48]

octadecylamine (ODA) BN [53]

amine-terminated polyethylene glycol BN [53]

N2H4, H2O2, HNO3/H2SO4, oleum BN [55]

Page 12: Graphene-analogous low-dimensional materials

11

Besides the use of polar solvents, liquid-phase exfoliation based on chemical

reactions is another facile route to isolate layered materials. This route, as a matter of

fact, is a combination of, firstly, a chemical modification of a layered material, and

later an ultrasonic treatment in organic solvent. As a first attempt, Lin et al. [53,54]

reported the covalent functionalization of h-BN by amine molecules through Lewis

acid-base reaction, in which the amine species forms covalent complexes with the

electron-deficient B atoms in h-BN surface. This complexation facilitates the

intercalation of functional molecules into the h-BN layered structures and triggers the

exfoliation of functionalized h-BN into solubilized thin nanosheets (3-20 layers) as a

result. In a later report, Nazarov et al. [55] reported the BN functionalizaiton with

inorganic reagents by heating h-BN power with hydrazine (N2H4), H2O2,

HNO3/H2SO4, and oleum. Further ultra-sonication in water or DMF leads to

exfoliation of the functionalized h-BN to form stable colloidal solutions of few-layer

BN nanosheets with lateral dimensions below 1 µm. Very recently, Sainsbury et al.

[ 56 ] have reported the surface covalent functionalization of exfoliated h-BN

nanosheets via nitriene addition reaction, and the generated product demonstrates

enhanced dispersability in organic solvents and facilitates the chemical

compatibilization of h-BN nanosheets within polymer matrices. Besides BN, this

method has also been extensively utilized to realize nanosheets of metal oxides and

hydroxides [57,58].

Another strategy used for exfoliation of layered materials is the ion-intercalation

method, which has been used a lot for exfoliating transition-metal dichalcogenides.

Page 13: Graphene-analogous low-dimensional materials

12

This method usually includes two basic steps: Li intercalation into the layered space

of bulk materials, and the subsequent exfoliation by immersing the Li-intercalated

compounds in water by ultrasonication. The reaction between the incorporated Li and

water forms LiOH and H2, which expands the interlayer space to facilitate the

exfoliation. At early attempts, single-layer MoS2, WS2, and their related

dichalcogenides were prepared by using butyl lithium as the intercalation reagent [59].

However, the major difficulty in this standard intercalation technique arises from the

low Li concentration, and more rigid conditions, such as elevated reaction

temperature (~100oC) and longer reaction time (~3 days), are generally required to

increase the Li content.

Zeng et al. [18] demonstrated a facile route for high-yield production of

single-layer dichalcogenides (MoS2, WS2, TiS2, TaS2, and ZrS2) through a

controllable electrochemical lithiation process (Fig. 2), which can be easily performed

within 6 hours at room temperature. The electrochemical lithiation was performed in a

battery test system, in which the layered materials were prepared as the cathode, and

the lithium foil was used as the anode. This method has the advantage that the degree

of lithium intercalation is effectively monitored by the discharge process. Subsequent

ultrasonication in water or ethanol yields large-amount single-layer nanosheets, and

the resulting MoS2 monolayers even achieve a high yield of 92%. The developed

electrochemical lithiation can be applied to prepare other 2D materials, like h-BN and

metal selenides or tellurides (such as NbSe2, WSe2, Sb2Se3, and Bi2Te3) [60].

Page 14: Graphene-analogous low-dimensional materials

13

Fig. 2. Scheme of electrochemical lithiation process. Reproduced from Ref. [18].

Copyright © 2011, Wiley-VCH.

A noteworthy drawback for the Li-intercalation method is that Li intercalation

results in a striking structural phase transformation [61]. Taken MoS2 as an example,

the Mo coordination changes from trigonal prismatic (2H) to octahedral (1T) after Li

insertion. The phase transformation process is detrimental since it leads to formation

of metastable MoS2 structures, and degrade their electronic behaviors and device

functionalizations.

The chemically derived solution processing, relying upon strong polar solvents,

reactive reagents, or ion intercalation, are versatile and up-scalable. The fabricated 2D

materials can be redispersed in common organic solvents, and can be used to deposit

in various environments and substrates not available for mechanical cleavage methods.

This method opens up a whole new range of potential large-scale preparations and

applications of 2D materials for nanodevices, composites, or liquid phase chemistry.

2.3 Chemical vapor deposition

In close resemblance to growth of graphene by thermal decomposition of

Page 15: Graphene-analogous low-dimensional materials

14

hydrocarbons on substrates, production of other 2D materials can also be

accomplished by high-temperature chemical reactions of molecular precursors on a

surface. This bottom-up self-assembly method is also known as chemical vapor

deposition (CVD). The surface not only serves as a template but also plays the role of

a catalyst to assist the epitaxial growth of solid films. The advantageous aspect for

this surface-assisted method is that clean single layer or few layers can grow without

accompanying thicker flakes.

The application of CVD technique in generating one-atom-thick BN films was

traced back to almost a decade ago. Generally, well-ordered single and double layer

BN domains were synthesized by thermal decomposition of borazine (BN)3H6 or

B-trichloroborazine (ClBNH)3 on catalytic metal surfaces such as Ni, Pd, Pt, Cu, Ag,

Fe, Ir, and Rh [62-64]. During these CVD processes, the rigid growth conditions, such

as high temperature (> 700oC) and ultrahigh vacuum chambers, are required. For the

deposited h-BN, B atom favors to site near a face-centered cubic (fcc) or hexagonal

close-packed (hcp) hollow site, and N atom prefers to locate close to the on-top

position of surface metal atoms [65].

These early studies on CVD growth only demonstrated the formation of BN

nanomeshs with relatively small areas, and the thickness, in most cases, is limited to

one layer. Methods for preparing h-BN layers with large areas and varying thickness

have been proposed in later reports. Recently, through atmospheric pressure (AP) or

low pressure (LP) thermal catalytic CVD methods, polycrystalline Ni film and Cu foil

have proved excellent catalytic substrates for growth of large-area thin BN nanosheets

Page 16: Graphene-analogous low-dimensional materials

15

(one to five layers) with borazine [ 66 ], amineborane [ 67 - 70 ], or decaborane

(diborane)/ammonia mixture [71,72] as the precursor. Besides metal substrates, BN

nanosheets with thickness lower than 5 nm can be directly fabricated on Si/SiO2

substrates by thermal CVD using B, MgO, and FeO powers as precursors under NH3

gas flow [73], or by catalyst-free microwave plasma (MP) reaction of BF3-H2-N2

mixtures [74]. In further advance, Liu et al. [75] reported the use of graphene as a

supported substrate for depositing BN nanosheets via a two-step CVD process (Fig. 3).

During the first step, graphene is deposited onto Cu foil by the chemical

decomposition of n-hexane at 950 C. During the second step, h-BN layers form on the

generated graphene/Cu foils through thermal decomposition of ammonia borane

(NH3-BH3) at 1000 C. The CVD-produced BN nanosheets can also be realized by

using other precursors (see Table 2). Besides BN, growth by CVD is a promising

approach to produce B-C-N hybrid materials [76].

Fig. 3. Schematic for preparation of graphene/BN stacked film. (a) Cu foil substrate.

(b) Graphene is grown on Cu. (c) Graphene/Cu foil is loaded for growth of h-BN film

on top. (d) and (e) are photographs of graphene (purple), graphene/h-BN film (blue),

and SiO2 (light purple). Reproduced from Ref. [75]. Copyright © 2011, American

Chemical Society.

Page 17: Graphene-analogous low-dimensional materials

16

Moreover, CVD methods have been used to grow atomically thin sheets of MoS2

on insulating substrates. For example, Zhan et al. [77] synthesized large-area MoS2

atomic layers (one to three layers) on SiO2/Si substrates by CVD using Mo and S as

reagents. In this procedure, the Mo thin films are firstly deposited onto SiO2

substrates by e-beam evaporation, and the pre-deposited Mo layers then react with the

introduced S vapors at 750 C to form MoS2 sheets (Fig. 4). As another alternative

routine, Lee et al. [78] fabricated MoS2 thin films directly deposited on SiO2/Si

substrates by CVD from MoO3 and S precursors under 650 C. Besides, CVD growth

of MoS2 was realized by using CVD-grown graphene on Cu foil as the template,

which gives rise to single crystalline hexagonal MoS2 flakes with a lateral size of

several micrometers [79].

Fig. 4. Schematic for growth of MoS2 on SiO2 substrate. Reproduced from Ref. [77].

Copyright © 2012, Wiley-VCH.

As disclosed by the above examples, surface-assisted CVD growth of thin layers

allows the patterning and large-scale growth of 2D materials, and the yielded

structures can be free of thicker nanosheets. However, this method relies upon the

Page 18: Graphene-analogous low-dimensional materials

17

interaction between substrate surface and the deposited films, and a very harsh

environment, such as high reaction temperature, is often required, which always poses

significant experimental difficulty. The representative examples are summarized in

Table 2 for 2D materials achieved by CVD techniques. Compared with the simple and

easily-handled chemical exfoliation methods, the CVD strategy is highly costly and

thus might limit its wide applications.

Table 2

Representative 2D materials produced by CVD method.

2D materials Precursor Method T ( C) Ref.

BN borazine APCVD 400 (1000) [66]

BN decaborane/ammonia thermal CVD 1000 [71]

BN diborane/ammonia LPCVD 1025 [72]

BN B, MgO, FeO, NH3 thermal CVD 900-1200 [73]

BN BF3/H2/N2 MPCVD 800 [74]

BN BCl3/NH3/H2/N2 thermal CVD 1000 [80]

BN boron oxide/melamine thermal CVD 1100-1300 [81]

BN (MeO)3B oxidation-nitrification

CVD 627 [82]

hybrid h-BNC methane/amineborane thermal CVD 900-1000 [76]

MoS2 Mo/S vapor thermal CVD 750 [77]

MoS2 MoO3/S thermal CVD 650 [78]

Page 19: Graphene-analogous low-dimensional materials

18

2.4 Surface-assisted epitaxial growth

Surface-assisted epitaxial growth can be regarded as a modification of CVD

method, in which the substrate surface serves as a seed crystal other than a template or

a catalyst. This method is alternatively considered as molecular beam epitaxy (MBE)

growth. The epitaxial growth has been successfully applied to fabricate

one-atom-thick Si sheets (silicene), and the un-reactive metal Ag with six-fold surface

symmetry provides a promising substrate to facilitate growth of hexagonal silicene or

Si nanoribbons [83].

On Ag(110) surface, in situ deposition of Si sources under untrahigh vaccum

conditions produces self-aligned Si NRs with a honeycomb graphene-like structure

[84-86]. For example, De Padova et al. [85] produced high aspect ratio Si NRs on

Ag(110) substrate, with several nanometers in lengths, 1.6 nm in width and only 0.2

nm in height. In a later work, De Padova et al. [86] grew perfectly straight and aligned

multilayer Si NRs with pyramidal cross section.

On Ag(111) surface, Lalmi et al. [87] synthesized a highly ordered Si monolayer

with a ( 3232 × )R30 superstructure. However, the Si-Si distance (0.19 ± 0.01 nm)

determined by Lalmi et al. is much shorter than the expected value of 0.21 nm as

computed from the ( 3232 × )R30 model. This makes the structure of silicene on

Ag(111) speculative. Lay et al. [88] have recently showed that the discrepancy

concerning the Si-Si separation in Lalmi et al.’s scanning tunneling microscopy (STM)

results can be explained by their misinterpretation of pure Ag(111) surface as the

strained silicene layer.

Page 20: Graphene-analogous low-dimensional materials

19

Fig. 5. (a) STM image of 2D Si layer on Ag(111)-(1×1). (b) Density functional theory

(DFT) results for silicene on Ag(111): top and side view of the fully relaxed atomic

geometries of the model for silicene on Ag(111) surface. Reproduced from Ref. [89].

Copyright © 2012, American Physical Society.

After that, five other groups independently reported experimental evidences of

ordered silicene phases on the same surface. For example, Vogt et al. [89] reported

epitaxial growth of silicene on Ag(111) substrate by directly depositing evaporated Si

onto Ag samples at temperatures between 220-260 C. The STM topograph (Fig. 5a)

shows a Si adlayer covering Ag(111) surface terraces with a honeycomb-like

appearance. Structurally, the Si atoms are located either above Ag atoms or between

Ag atoms, and the silicene layer forms a uniform (4×4) symmetry with respect to the

unrestructured (1×1) Ag(111) surface. At variance with the planar skeleton of

graphene, silicene prefers to be structurally low-corrugated. The in-plane Si-Si

distance determined from the STM image is about 0.22 nm (±0.01 nm), in agreement

with theoretical prediction of 0.232 nm (Fig. 5b). Similarly, Chiappe et al. [90], Feng

et al. [91], Lin et al. [92], and Jamgotchian et al. [93] also produced silicene

monolayer on Ag(111) substrate. According to their results, the silicene sheets are

rather sensitive to deposition temperature, and the resulting silicene domains present

Page 21: Graphene-analogous low-dimensional materials

20

different orientations ((4×4), ( 1313 × ), ( 77 × ), ( 3232 × )) with respect to the

(1×1) Ag substrate on varying temperatures.

Besides the Ag substrate, silicene can also be fabricated on Ir(111) surface [94].

Moreover, Fleurence et al. [95] showed that silicene can form through surface

segregation on ZrB2 (0001) thin films grown on Si(111) wafers.

2.5 Synthesis of 1D nanoribbons

1D nanoribbons have properties distinctive from their 2D nanosheets due to edge

states. Growth of high-quality nanoribbons with controllable widths and smooth edges

is important for many technological applications. However, compared with numerous

routes established for fabricating graphene nanoribbons, the experimental

availabilities for facile synthesis of other nanoribbons are comparatively rare. Besides

the availability of silicene nanoribbons, BN, MoS2 and WS2 nanoribbons have been

successfully fabricated by unique experimental techniques.

2.5.1 Unzipping BN nanotubes to produce BN nanoribbons

In analogy to the successful unzipping of carbon nanotubes to fabricate graphene

nanoribbons, BN nanoribbons have been successfully produced by unzipping of BN

nanotubes. Similarly, multiwalled BN nanotubes were unzipped to form nanoribbons

by selective plasma etching or by potassium vapor intercalation.

Initially, Zeng et al. [19] fabricated BN nanoribbons based on Ar plasma etching

of BN nanotubes embedded within a polymer (Fig. 6a). In this process, since side and

bottom walls of the nanotubes were initially imbedded in the poly(methyl

Page 22: Graphene-analogous low-dimensional materials

21

methacrylate) (PMMA) film, only the top part of the tube-shell was exclusively

etched and gradually removed, resulting in nanotube unzipping through cutting top

walls. The produced BN nanoribbons mainly displayed N-terminated zigzag edges

and surface vacancy defects, with the ribbon’s widths as narrow as 15 nm and the

lengths ranging from several hundred nanometers to several microns.

Erickson et al. [20] developed another synthetic method to produce BN

nanoribbons by the K-intercalation-induced longitudinal splitting of BN nanotubes

(Fig. 6b). Mechanically, K intercalation between the walls induces bond strains

circumferentially around the sidewalls of the tubes, which facilitates bond breakage

along the longitudinal direction of outer walls. This approach results in formation of

narrow (20-50 nm), long (at least 1 µm in length), thin (usually between 2 and 10

layers) and high crystalline BN nanoribbons.

Very recently, Li et al.[96] have reported an in situ unzipping of BN nanotubes

to produce BN nanoribbons, where the unzipping arises from intercalation of BN

nanotubes by Li-NH3 species formed during nanotube synthesis.

Fig. 6. (a) Schematic of the unwrapping process of BN nanotubes induced by plasma

etching and atomic force microscopy (AFM) image of single-layered BN nanoribbons.

Page 23: Graphene-analogous low-dimensional materials

22

Reproduced from Ref. [19]. Copyright © 2010, American Chemical Society. (b)

Schematic of the potassium-intercalation-induced splitting process of BN nanotubes

and TEM micrograph of a few-layer BN nanoribbon splitting off its parent ribbons.

Reproduced from Ref. [20]. Copyright © 2011, American Chemical Society.

2.5.2 Growth of ultranarrow MoS2 and WS2 nanoribbons inside carbon nanotubes

One method currently developed for producing ultra narrow MoS2 and WS2

nanoribbons is the direct chemical growth inside carbon nanotubes (Fig. 7a). This

method was developed by Wang et al. [21], and the carbon nanotubes behave as the

role of templates that confine the growth of nanoribbons along 1D direction. In this

synthetic method, H3PMo12O40 was firstly encapsulated into carbon nanotubes, and

the filled carbon nanotubes were then heated under H2S/H2 to generate MoS2.

Depending on the diameter of carbon nanotubes, single-layered and bi- or tri-layered

MoS2 nanoribbons can be obtained. Fig. 7b shows the HRTEM image of a

single-layer MoS2 nanoribbon encapsulated in a single-walled carbon nanotube. The

MoS2 nanoribbons clearly exhibit a zigzag-typed edge with four zigzag chains along

the ribbon width (Fig. 7c is the schematic model). Through similar procedure, ultra

narrow WS2 nanoribbons with smooth zigzag edges and 1-3 layered thickness were

fabricated inside carbon nanotubes using H3PW12O40 as the starting material [22].

This method can be potentially applied to yield other transition-metal chalcogenide

nanoribbons.

Page 24: Graphene-analogous low-dimensional materials

23

Fig. 7. (a) Model of MoS2 nanoribbon inside carbon nanotube. (b) HRTEM image of

a single-layer MoS2 nanoribbon encapsulated in a single-walled carbon nanotube. (c)

Model of the MoS2 nanoribbon in (b). Reproduced from Ref. [21]. Copyright © 2010,

American Chemical Society.

2.6 Other synthetic routes

The above reviewed methods are the typically adopted methods for synthesizing

2D materials, especially for BN and dichalcogenides. Besides, many other routes have

been developed for fabricating thin sheets of BN, metal dichalcogenides and other 2D

materials (see Table 3). For example, BN nanosheets can be realized by ex-situ ion

etching [97], surface segregation on catalytic metals [98], and high-energy electron

beam irradiation inside TEM [99,100]. Wang et al. introduced a “chemical blowing”

method to synthesize BN and C-containing BN nanosheets (one to few layers) in mass

production [101]. This method relies on making large bubbles with thin B–N–H (or

B–N–C–H) polymer walls by releasing H2 from a precursor ammonia borane, in

which the produced H2 bubbles promote thinning process of BN nanosheets.

Moreover, MoS2 nanosheets can be alternatively achieved by laser-thinning technique

Page 25: Graphene-analogous low-dimensional materials

24

[ 102 ], thermolysis of ammonium thiomolybdate ((NH4)2MoS4) [ 103 ], thermal

evaporation technique [104], and hydrothermal synthesis [105], etc. Thin nanosheets

of other metal dichalcogenides/chalcogenides, such as VS2 [106], MoSe2, WSe2 [51],

TiS2 [107], Bi2Se3 [108], and Bi2Te3 [109], have also been fabricated through various

experimental skills.

Table 3

Other synthetic routes for synthesizing BN and metal dichalcogenides or

chalcogenides.

2D materials Synthesis method Ref.

BN ex-situ ion etching [97]

BN surface segregation from bulk Fe-Cr-Ni alloy doped with B and N [98]

BN chemical reaction of boric acid with urea at 900oC [110]

BCN chemical reaction of activated charcoal with boric acid and urea at

900oC [111]

BN high-energy electron beam irradiation [99,100]

BN “chemical blowing” method using ammonia borane as precursor [101,112]

MoS2 laser-thinning of multilayered MoS2 flakes [102]

MoS2 two-step thermolysis of (NH4)2MoS4 precursor [103]

MoS2 thermal evaporation of MoO3 with S powers [104]

MoS2 annealing of ball-milled MoO3 and S powers at 600oC [113]

MoS2 solvothermal synthesis from MoO3 and thioacetamide [114]

WS2 sulfidation of W18O49 nanorods [115]

MoS2 hydrothermal reaction between MoO3 and KSCN [105]

MoSe2 hydrothermal reaction between molybdic acid and Se [51]

MoS2, WS2 chemical reaction of molybdic (tungstic) acid and thiourea at 773K [105]

MoSe2, WSe2 chemical reaction of molybdic (tungstic) acid and Se at 773K [51]

VS2 chemical exfoliation of VS2

.NH3 precursor formed by hydrothermal reaction of NH4VO3 and thioacetamide

[106]

Page 26: Graphene-analogous low-dimensional materials

25

TiS2 chemical reaction of TiCl4 and oleylamine/S mixtures at 300oC [107]

Bi2Se3 molecular beam epitaxy (MBE) growth [108]

Bi2Te3 surface-assisted chemical vapor transport (CVT) technique [109]

In addition, the unique fabrications for graphitic-like ZnO and MXene layers, as

well as coordination synthesis for metal organic frameworks (MOFs) and

polymerization reaction for assembling covalent organic frameworks (COFs) will be

reviewed in their corresponding sections later.

2.7 Characterization

The layer thickness of 2D materials is determined mainly by optical microscopy

imaging, AFM, and Raman spectroscopy [116]. A facile observation of the obtained

2D sheets can be realized by optical microscopy, and further combination with AFM

technique offers a fast estimation of the thickness distribution. Additionally, since

vibrational spectrum shows significant thickness dependence, Raman spectroscopy is

also widely used to determine the thickness and to examine the changes in material

properties with thickness (Fig. 8). For example, the AFM images in Fig. 8a and 8d

illustrate the identification of regions with different thicknesses (one to four layers)

for extracted BN and MoS2 nanosheets deposited on oxidized Si wafer. With regard to

Raman spectroscopy, the Raman intensity and the peak shift were employed for

characterizing layer numbers. The characteristic BN peak centered at ~1366 cm-1 in

bulk h-BN originates from the E2g mode of B-N vibration within h-BN layers, which

would shift to higher or lower frequency under compressive or tensile stress. The

monolayer h-BN exhibits a blue shift of up to 4 cm-1, which is explained by hardening

Page 27: Graphene-analogous low-dimensional materials

26

of the E2g mode because of the slightly shorter B-N bonds in BN monolayer. In

bilayer h-BN, the strain effect becomes dominated, and causes the red shift by 1-2

cm-1. The peak becomes stronger as the thickness increases, ascribed to a low yield of

monolayer BN compared with thicker ones (Fig. 8b). Similarly, for the exfoliated

MoS2 layers, the in-plane E2g1 and the out-of-plane A1g vibration modes near 400 cm-1

are sensitive to film thickness; the former shows red shift and the latter blue shift as

thickness increases, and finally the Raman frequencies converge to bulk values when

the films are thicker than four layers (Fig. 8f).

Fig. 8. (a) Optical microscope of thin BN nanosheets. The insets show AFM images.

(b) Raman spectroscopy of thin BN nanosheets. Reproduced from Ref. [32].

Copyright © 2011, Wiley-VCH. (c) Optical micrograph of thin MoS2 films. (d) AFM

image of MoS2. (e) Model of MoS2 bilayer structure. (f) Raman spectra of thin and

bulk MoS2 films. (g) Atomic displacements of bulk MoS2 crystal. Reproduced from

Ref. [33]. Copyright © 2010, American Chemical Society.

Page 28: Graphene-analogous low-dimensional materials

27

2.8 Theoretical methods

The advancement of parallel computational powers, especially development of

supercomputer facilities in recent decades, has rapidly paced the theoretical

investigations on various materials ranging from bulk to low-dimensional

nanostructures. The significance of theoretical studies is certainly self-evident, which

can not only provide scientific understanding to elucidate experimentally observed

phenomena, but also offer an important tool for conducting materials design and

property prediction. The fundamental theory for materials modeling and computations

at the atomic scale is based on ab initio or quantum mechanics theory.

Density functional theory (DFT) is currently the most commonly employed

quantum mechanics method, which has evolved into a powerful tool for computing

electronic ground-state properties of a large number of nanomaterials. The entire field

of DFT method relies on the theorem that the ground-state energy of a many-electron

system is a unique and variational functional of the electron density, and this

conceptual proposal is implemented in a mathematical form to solve the Kohn-Sham

(KS) equations.

Within the DFT frameworks, the choices for computational levels and basis sets

vary over a large range. Typically, spin-polarized or non-polarized computations can

be performed with the exchange correlation functional based on a local density

approximation (LDA) considering Perdew-Wang correlational (PWC) [ 117 ], a

generalized gradient approximation (GGA) considering Perdew-Wang (PW91) [118]

or Perdew-Burke-Eruzerhof (PBE) functionals [119], or hybrid functionals (such as

Page 29: Graphene-analogous low-dimensional materials

28

HSE [120]) taking into account the non-local exchange correlations. The basis sets

widely employed include double numerical basis sets plus p-polarization functions

(DNP) or d-polarization functions (DND) [121,122], double-zeta plus polarization

atomic-orbital basis sets (DZP) [ 123 , 124 ], and plane-wave basis sets [ 125 ].

Particularly, the plane-wave DFT computations are more regularly adopted for

studying nanomaterials [126]. Moreover, to reduce the computational burden caused

by core electrons, psuedopotentials based on the frozen core approximations are

always employed, which consist of effective core potentials [127], DFT-based

semi-core pseudopotentials (DSPP) [128], Troullier-Martins (TM) norm-conserving

pseudopotentials [129], Vanderbilt ultrasoft pseudopotentials (USPP) [130], and

projector-augmented-wave (PAW) pseudopotentials [126].

The computational programs that have been used mainly include VASP [125],

CASTEP [131], PWSCF [132], SIESTA [124], ABINIT [133], and DMol3 [122].

Among them, VASP, CASTEP, PWSCF, and ABINIT are plane-wave codes using

plane-wave basis set and norm-conserving or ultrasoft pseudopotentials, or PAW

method, while DMol3 and SIESTA are all-electron codes using numerical

atomic-orbital basis sets and all-electron or semicore or effective-core

pseudopotentials (for DMol3) and standard norm-conserving pseudopotential (for

SIESTA).

Note that, although DFT computations give powerful predictions for a wide

range of materials and properties, there are still some situations in which the standard

DFT is not physically accurate. For example, both LDA and GGA functionals are

Page 30: Graphene-analogous low-dimensional materials

29

well known to underestimate band gaps of semiconductors, with errors commonly

larger than 1 eV compared with experimental values. In particular, the self-interaction

error associated with the classical DFT often yields a qualitatively incorrect prediction

for the band gaps and magnetic properties of strongly correlated systems with robustly

localized electrons. Moreover, the standard DFT is only restricted to evaluate the

ground-state energy and properties, and can not deal with computations of electronic

excited states. Also, the standard DFT computations can not well describe long-range

weak interactions, in which LDA tends to overestimate the binding strength, while

GGA tends to behave opposite.

To overcome these inherent limitations, significant corrections have been

introduced to improve the accuracy of DFT computations. Taken as examples,

applying GW approximation to the electron self-energy is successful in accurately

predicting the excited-state properties and improving on the Kohn-Sham band

structure of semiconductors [134]. The screened exchange hybrid density functional,

HSE, has also been successful in predicting the band gaps of semiconductors, and it

appears to provide a good starting point for perturbative GW corrections. Besides, the

DFT+U method [135], which adds a model Hubbard-U term to the simple DFT

functional, provides a qualitatively correct treatment of the electronic structures of

strongly correlated materials. Moreover, the DFT+D method [136], which introduces

dispersion corrections to the conventional Kohn-Sham DFT energy, has been

developed to remedy drawbacks of DFT regarding the dispersion forces.

In the last paragraph of this section, we will briefly sum up the structural

Page 31: Graphene-analogous low-dimensional materials

30

modeling of 2D nanosheets and 1D nanoribbons. The 2D sheets under investigation

can be constructed directly from layered bulk materials, or are assembled based on

experimental measurements. Depending on the cutting directions, two types of

nanoribbons with either zigzag or armchair shaped edges can be generated. Following

the conventional notation of graphene nanoribbons, the graphene-analogous

nanoribbons are also specified by the ribbon parameter Nz or Na, which is defined as

the number of parallel zigzag chains or the number of dimer lines along the width

directions of zigzag or armchair ribbons, respectively. Taking BN as an example, Fig.

9 clearly illustrates geometric structures of 2D BN monolayer together with its cutted

zigzag and armchair BN nanoribbons with 9 zigzag chains (9-zigzag BNNR) and 15

dimer lines (15-armchair BNNR), respectively. Other types of graphene-analogous

materials are modeled in a similar way.

Fig. 9. BN monolayer (a) and its derived 9-zigzag BNNR and 15-armchair BNNR (b).

3. Planar graphene analogues

In the following sections, we will devote our attention to atomic structures,

material properties, and potential applications of various 2D and 1D materials beyond

graphene and graphene nanoribbons, which are divided into planar graphene

Page 32: Graphene-analogous low-dimensional materials

31

analogues (Section 3), hypothetical planar materials that have not been experimentally

realized (Section 4), non-planar materials (Section 5), as well as coordination and

covalent organic polymers (Section 6).

3.1 “White graphene”: BN nanosheets and nanoribbons

3.1.1 Comparison between BN and C

BN nanomaterials deserve special attention since they are carbon’s isoelectronic

analogues. In fact, BN exhibits various crystalline forms in many ways analogous to

carbon, including diamond-like cubic BN, graphite-like h-BN, wurtzite-like BN,

onion-like fullerenes, and BN nanotubes. Within these polymorphs, h-BN is

thermodynamically the most stable and softest form, and has attracted increasing

attention. Bulk h-BN has a similar layered structure as graphite with the exception

that the basal planes in h-BN are positioned directly on top of each other, with the

electron-deficient B atoms in one layer lying over and above the electron-rich N

atoms in adjacent layers (Fig. 10). For graphite, however, the adjacent layers are

stacked offset, and alternating C atoms lie above and beneath the hexagon centers.

Both bulk h-BN and graphite exhibit very similar lattice constants and interlayer

distances (lattice constants: a = 2.456 Å, c = 6.696 Å for graphite, and a = 2.504 Å, c

= 6.661 Å for bulk h-BN; interlayer distances: 3.33-3.35 Å for graphite, and 3.30-3.33

Å for h-BN). Due to their close structural similarity, monolayer h-BN can thus be

regarded as a structural analogue of graphene where the C-C bonds are substituted by

alternating B-N pairs. In analogy to the common use of “white graphite” for bulk

Page 33: Graphene-analogous low-dimensional materials

32

h-BN, the monolayer h-BN is depicted as “white graphene” [19].

Fig. 10. Graphitic structures for carbon and h-BN.

In spite of their structural resemblance, h-BN nanosystems show strikingly

distinct properties from carbon counterparts. The difference in electronegativity

between boron and nitrogen induces ionicity, which is responsible for narrowing of

sp2 derived π bands and the corresponding loss of conductivity, causing h-BN, in

contrast to semimetallic graphite, to be highly insulating. As a matter of fact, h-BN

based materials, either 3D bulk, 2D sheets, or 1D nanotubes, are all electrically

insulating with wide band gaps of 5-6 eV.

On the other hand, h-BN systems also demonstrate advantageous properties. For

example, BN materials are thermally and chemically more stable than carbon

counterparts, and possess a superb thermal conductivity, extraordinary mechanical

robustness, excellent resistance to oxidation (until temperature over 800 C), and good

optical properties. The high thermal stability of h-BN is given by the strong covalent

Page 34: Graphene-analogous low-dimensional materials

33

network in the plane and the strong polarizability of individual layers. Note that due

to their oxidation resistance, thin h-BN nanosheets can not be prepared by the

oxidation/exfoliation methods as widely adopted for large-quantity fabrication of

graphene.

3.1.2 Electronic and magnetic properties of BN nanosheets and nanoribbons

The isolated BN monolayer inherits the insulating characteristic of bulk h-BN,

and exhibits an indirect band gap (4.3 eV for LDA and 6.0 eV for GW correction)

[137]; whereas the experimentally measured band gap is 5.97 eV [138], larger than

that of the bulk h-BN (5.2-5.4 eV). For 1D BNNRs, since quantum size and symmetry

effects as well as edge effects all emerge, the electronic and magnetic properties of

BNNRs differ a lot from those of BN sheets, which are closely associated with edge

geometries and depend critically on how the edges are passivated, especially for

zigzag edges [139].

Similar to GNRs, spin-polarized magnetic behaviors were only found in

zigzag-typed BNNRs [140]. Armchair-typed BNNRs, owing to strong coupling of the

dangling bonds of dimeric N and B atoms at the same edge, are ground-state

non-magnetic and exhibit semiconducting behaviors with large indirect gaps [141].

The lowest-energy magnetic configuration of bare zigzag BNNRs corresponds to a

ferromagnetic (FM) spin arrangement at the N edge and an anti-FM spin arrangement

at the B edge (corresponding to (+ , ++) magnetic configuration, Fig. 11). This

behavior is in contrast with that found in zigzag GNRs, where both zigzag C edges

Page 35: Graphene-analogous low-dimensional materials

34

interact via anti-FM spin arrangement mediated by the carbon backbone. Note that

HSE functional predicts all the five spin configurations of bare zigzag BNNRs to be

semiconducting, whereas semilocal PBE functional predicts a half-metallic behavior

for them.

Fig. 11. Five possible magnetic configurations for bare zigzag BN nanoribbons.

Reproduced from Ref. [140]. Copyright © 2008, American Chemical Society.

The bare zigzag edges of BNNRs are less stable than the armchair ones due to

the existence of un-paired edge dangling bonds, and also have less stability than the

reconstructed edge made of 5-7 membered rings [142], which, however, are stabilized

by passivating the dangling edges with H atoms. When only one edge is passivated,

the zigzag BNNRs with passivated B edge and bare N edge are ferromagnetic half

metals [143]. Conversely, the zigzag BNNRs with bare B edge and passivated N edge

are antiferromagnetic p-type semiconductors with an indirect band gap [25,144].

When both edges are hydrogen passivated, the fully passivated armchair and

zigzag BNNRs are all non-magnetic semiconductors with direct and indirect band

gaps, respectively, and their band gaps are strongly affected by external electric field

(Fig. 12) [145]. Band gaps of armchair BNNRs exhibit a family oscillation behavior

Page 36: Graphene-analogous low-dimensional materials

35

as width increases, and then converge to a value about 0.02 eV smaller than the BN

monolayer (4.53 eV) due to the presence of weak edge states. Under a transverse

electric field, the band gaps of armchair BNNRs are reduced with increased field

strength regardless of the field direction. In contrast to armchair BNNRs, band gaps of

zigzag BNNRs monotonically decrease with widths and converge to a value about 0.7

eV smaller than the BN monolayer due to the presence of strong edge states.

Furthermore, the polarized edge states make zigzag BNNRs exhibit an asymmetric

response to the electric field.

Fig. 12. (a) LDA band gaps of BNNRs versus widths, and dashed lines indicate the

band gap of BN monolayer. (b) Band gaps of 36-armchair BNNR (filled red circles)

and 84-armchair BNNR (empty blue squares) versus field strength. (c) Band gaps of

27-zigzag BNNR (filled red circles) and 46-zigzag BNNR (empty blue squares)

versus field strength. Reproduced from Ref. [145]. Copyright © 2008, American

Chemical Society.

Page 37: Graphene-analogous low-dimensional materials

36

In the above considered BNNRs, the edge B (N) atoms are of sp2 type. Another

attractive case is the bihydrogenated edge under hydrogen rich environment, where

the fully saturated edge B (N), bonded to two hydrogen atoms, is of sp3 type [146].

Specifically, zigzag BNNRs composed of bihydrogenated B edge and

monohydrogenated or bihydrogenated N edge are all ferromagnetic metals. Armchair

BNNRs, however, are robust non-magnetic semiconductors regardless of hydrogen

contents.

3.1.3 Band-gap modifications of BN nanosheets and nanoribbons

As mentioned in the previous section, perfect BN nanosheets always have wide

band gaps, which are far beyond the gap values (< 3 eV) desired for most of the

current electronic and optical devices, thus become a severe obstacle in processing

BN-based electronics. The well-adopted strategies, such as defects, doping, as well as

surface and edge modifications, can effectively modulate the electronic properties of

BN monolayer and nanoribbons and enhance their device applicability.

3.1.3.1 Defects

The structural and electromagnetic properties of BN nanosheets/ribbons with

unperturbed high perfection of atomic lattices are outstanding; however, structural

defects, in the forms of intrinsic point defects and extrinsic substitution or doping,

might be present during growth or processing. Deviations from perfections, however,

can be useful in some applications since they make it possible to tailor local properties

of BN lattice and to achieve new functionalities. In this section, we will present

Page 38: Graphene-analogous low-dimensional materials

37

experimental and theoretical surveys on structural defects of BN nanosheets/ribbons,

and discuss influence of these defects on the properties of BN.

Point defects. The commonly observed point defects in experimentally

synthesized BN sheets are triangle vacancy defects with missing B and N atoms.

These vacancies with different sizes have been detected by HRTEM and can in

principle be created in a controlled manner by electron beam irradiation onto BN

sheets (Fig. 13). Jin et al. [99] and Meyer et al. [100] have fabricated the defective

single layer h-BN nanosheets at the beam energy of 120 kV and 80 kV inside the

TEM. Mono-atomic and multi-atomic vacancies, especially B monovacancy and

N-terminated triangle multivacancies, dominated in these BN sheets [147]. The

triangle-shaped vacancies are created mainly due to knock-on effect, which is

mechanically a quasielastic collision between incident electron and nuclei of the

atoms belonging to the h-BN specimens. Since the defects are created under a

high-energy electron beam, the formation and evolution of the triangle-shaped

multivacancies are governed mainly by kinetic factors such as knock-on displacement

rates or probabilities. The preference of B vacancy indicates that boron atoms are

more easily removed than nitrogen atoms.

Theoretical computations showed that structural stabilities of triangle vacancies

depend strongly on the environmental conditions [148-150]. Under nitrogen-rich or

electron-rich conditions, triangle vacancies with N-terminated edges (such as VB and

VB+3N) are more stable. In contrast, under boron-rich conditions, B-terminated edges

(such as VN and VN+3B) are energetically more favorable. The BN diatomic vacancy

Page 39: Graphene-analogous low-dimensional materials

38

VBN is stable under nitrogen-rich and neutral conditions.

Fig. 13. (a) HRTEM showing lattice defects in h-BN. (b) Models for atomic defects in

h-BN. Reproduced from Ref. [99]. Copyright © 2009, American Physical Society.

The point defects observed in BN sheets are markedly different from those in

graphene. In graphene, an energetically favorable defect is a five-membered ring

(5MR), pentagon. As a consequence, topological defects like reconstructed

pentagon-heptagon or other odd-numbered rings are readily generated in graphene

under electron beam irradiation, especially at edges. However, no reconstructed

vacancy (5MRs) or Stone-Wales (SW) defect (pentagon-heptagon defect) [151] was

observed in h-BN sheets. This is because formation of homonuclear N-N bonds or

B-B bonds is highly unfavorable, making the local reconstruction to form pentagon or

SW defects extremely difficult. As a comparison, Fig. 14 shows the HRTEM images

of BN and graphene membrances forming under electron beam irradiation at 80 keV.

It is clearly seen that a number of triangle-shaped monovacancies are observable,

which do not exhibit any structural reconstruction. While for graphene, three types of

Page 40: Graphene-analogous low-dimensional materials

39

defect structures are observed: a carbon monovacancy without reconstruction, a

reconstructed monovacancy forming a pentagon-nonagon (5-9) pair, and a

reconstructed pentagon-heptagon (5-7) SW defect. Notably, the reconstructed BN

divacancy (VBN, a Schottky defect pair) has been observed in BN nanotubes, but no

stable BN divacancy was identified in planar BN nanosheets.

Fig. 14. HRTEM images of (a) h-BN and (b) graphene membrane. Reproduced from

Ref. [100]. Copyright © 2009, American Chemical Society.

The electronic and magnetic properties of BN sheets [148,152] and ribbons [153]

with point defects have been investigated theoretically. Depending on the sizes and

edge terminations of created vacancies, BN monolayer should display substantial

magnetism in proportional to the number of unpaired electrons of the N/B atoms

along the zigzag edges. Especially, a single B vacancy in BN monolayer realizes a

half-metallic feature. Vacancy defects also induce spontaneous magnetization and

manipulate electronic properties of BN nanoribbons. The formation of vacancy occurs

preferentially at the vicinity of B edge than other sites. The effect of B or N vacancy

on zigzag BNNRs shows remarkable dependence on defect sites and concentrations,

Page 41: Graphene-analogous low-dimensional materials

40

and some special band structures such as spin-polarized semiconductor, half metal,

and spin-gapless semiconductor can be achieved. As a further addition of point

defects, line defects are also present [154]. The defect lines are located at the BN

domain boundaries which have different orientations. The existence of extended line

defects presents a new pathway to tune the properties of BN nanoribbons [155].

3.1.3.2 Substitution and adsorption doping

Foreign atoms (molecules) can be incorporated into the BN lattice or embedded

in the vacancy sites as substitutional impurities, or be adsorbed on the BN surface as

adsorption dopants [156]. Theoretical studies dealing with functionalization of BN

layer predicted that adsorption or substitution doping by magnetic metal impurities

(such as transition metals, V, Cr, Fe, Co, Cu, and Mn [157], and noble metals, Au, Ag,

Pt, and Pd [158]), and non-magnetic impurities (such as Be, B, C, N, O, Al, and Si

[ 159 ]) induce magnetism due to emergence of spin-polarized impurity states.

Localized impurity states occurring in the band gap yields interesting properties for

electronic, magnetic and optical applications.

Carbon doping, due to its atomic similarity with B and N, has attracted a lot of

attention. C-substitution for non-edge boron or nitrogen atom in BNNRs induces

spontaneous magnetization, which is independent on the site of substitution or type of

BNNRs [ 160 ]. Post-synthesis of substitutional C doping of h-BN sheets was

implemented via in situ electron beam irradiation inside an energy-filtering 300 kV

HRTEM (Fig. 15a) [161]. Using essentially the same technique, Krivanek et al. [162]

Page 42: Graphene-analogous low-dimensional materials

41

reported the substitution of both carbon (C substituting for B and N) and oxygen (O

substituting for N) atoms into the BN honeycomb network at electron energies of 60

keV (Fig. 15b). The substitutional C doping transforms BN nanostructures from

electrical insulators to conductors. The doping mechanism was proposed to rely on

the knockout ejections of B and N atoms and subsequent healing of vacancies with

supplied C atoms. Using DFT static and dynamic computations, Berseneva et al. [163]

explained that the C-substitution process is governed not only by the response of such

systems to irradiation, but also by the energetics of the atomic configurations, and it

costs less energy to substitute B atoms than N, especially when the system is

positively charged.

Fig. 15. (a) Schematic for electron-beam-induced substitutional C doping in a

honeycomb BN lattice. Reproduced from Ref. [161]. Copyright © 2011, American

Chemical Society. (b) DFT simulation of a BN layer containing the observed

substitutional impurities. Red, B; yellow, C; green, N; blue, O. Reproduced from Ref.

[162]. Copyright © 2010, Nature Publishing Group.

Drilling holes with an atomic-size electron beam and filling them with atoms

(such as C or other elements) was a promising method for constructing hybrid

Page 43: Graphene-analogous low-dimensional materials

42

structures with atomic precision and for tuning electronic properties of diverse B-C-N

structures in a full compositional range between pure BN and C systems.

Surface doping by organic molecules through non-covalent interaction is a

simple and effective method to tune electronic structure of BN nanosystems. Because

of interfacial charge transfer, surface adsorption with strong electron acceptor (tetra

cyanoquinodimethane, TCNQ) or electron donor (tetrathiafulvalene, TTF) molecule

can significantly reduce the wide band gap of pristine BN nanosheets/ribbons and

result in a p- or n-type semiconductor, respectively [164]. The origins for band

modifications are attributed to the introduction of new impurity states near the top

valence band or the bottom conduction band of pristine BN nanosheets/ribbons,

contributed by the highest occupied molecular orbital (HOMO) or lowest unoccupied

molecular orbital (LUMO) levels of the adsorbed molecules.

3.1.3.3 Surface modifications of BN nanosheets

The surface of BN nanosheets can be chemically decorated by various groups

such as hydrogen, fluorine and oxygen, and the hybridized states of B or N atoms in

these cases change from sp2 into sp3. Nevertheless, there is some difference in surface

functionalization between BN and graphene. Structurally, hydrogenated graphene

prefers chair-like configuration (graphane). For hydrogenated BN sheets, however,

the chair configuration is less favorable than boat or stirrup configurations [165]. Note

that the chair-like structure can be stabilized by surface modifications with groups of

large atomic radius such as F, Cl, OH, CH3, and NH2 [ 166 ]. Interestingly,

Page 44: Graphene-analogous low-dimensional materials

43

hydrogenation opens a band gap in graphene, but it reduces the band gap of BN sheets

[167]. At the DFT-GGA level, the hydrogenated BN monolayer has a direct band gap

of 3.33 eV. The semihydrogenated BN sheets can form by applying electric fields to

the fully hydrogenated ones or by single-side hydrogenation on the substrate. The

semihydrogenation on BN sheet is slightly different from graphene because of the

heteroatomic inequivalence of B and N atoms. Boron-semihydrogenated BN sheet is

more stable and behaves as a ferromagnetic half-metal [167].

Due to its strong electron affinity, F atom has a strong site preference to bond

with B, but a weak bonding with N. The magnetism and band structures in fluorinated

BN can be controlled by applying strain [168]. Promoted by the selectivity of

fluorination with B atoms, surface fluorination of few-layered BN sheets produces

stable F-terminated cubic BN nanofilms [169], which exhibit controllable band gap by

altering thickness or applying electric fields.

Moreover, in analogy to oxidation of graphene to form graphene oxide, the

surface oxidation of BN sheet generates the BN derivative, BN oxide, which tends to

form an O domain or O chain on the BN sheet, and introduces unoccupied impurity

states leading to band gap reduction of BN sheet [170].

Besides BN sheet, hydrogenation also plays a crucial role in engineering

electromagnetic properties of BNNRs, as indicated by Chen et al. [26]. The fully

hydrogenated armchair BNNRs act as non-magnetic semiconductors, while the fully

hydrogenated zigzag BNNRs behave as ferromagnetic metals. The partially

hydrogenated zigzag BNNRs exhibit diverse properties, where as hydrogenation

Page 45: Graphene-analogous low-dimensional materials

44

coverage increases, a semiconductor→half-metal→metal transition occurs,

accompanied by a nonmagnetic→magnetic transformation (Fig. 16). For all the

hydrogenated zigzag BNNRs, the bands around the Fermi level are dominated by the

edge B/N atoms with sp3 hybridization and the related H atoms, and increasing

hydrogenation ratio causes the highest valence bands and the lowest conduction bands

to shift towards the Fermi level and cross it eventually. Thus, controlling the

hydrogenation ratio can precisely modulate the electronic and magnetic properties of

zigzag BNNRs, which endows BN nanomaterials many potential applications in the

novel integrated functional nano-devices.

Fig. 16. Structural and electronic properties of partially hydrogenated 8-zigzag

BNNRs with different ratios of LH to L0: (a) 1:4, (b) 2:3, (c) 3:2, and (d) 4:1. (e)

Structure of fully hydrogenated zigzag BNNR. Parts f and g show zooms on the

region about the Fermi level of band structure and density of states (DOS) of (d).

Reproduced from Ref. [26]. Copyright © 2010, American Chemical Society.

Page 46: Graphene-analogous low-dimensional materials

45

It is worth noting that although completely and partially hydrogenated or

fluorinated graphene has been successfully fabricated, the surface-hydrogenated or

-fluorinated BN sheets are rarely available. So far, only the partially hydrogenated

few-layered BN membranes have been experimentally achieved by hydrogen plasma

treatment [171]. Moreover, the surface oxidation under strong oxidizing reagents is a

common reaction of graphene, but the oxidation reaction of BN counterpart becomes

extremely difficult due to its chemical inertness. With these arisen challenges, the

realizations of hydride-, fluoride- and oxide-functionalized BN derivatives are

expected to find their unique technical solutions in future.

3.1.3.4 Edge modifications of BN nanoribbons

The edges are typically more reactive than the surfaces, and edge modifications

can effectively control the electronic properties of BNNRs. Hydrogen atoms are the

typical modifiers for BN edges, and are usually used as the reference for comparison

with other edge modifiers. Chemical functionalization of zigzag BNNRs edges by Cl,

OH, and NO2 results in a narrowed band gap, while decorating with F atom leads to

an enlarged gap [172]. For these edge decorations, B edges are more favorable than N

edges for adsorption. Especially, edge functionalizations by F and OH give higher

stability compared with the hydrogen passivated counterparts, suggesting that these

modifications have great potentials to be realized experimentally. Interestingly, both

O-edge and S-edge terminated zigzag BNNRs become metallic [173]. The O-edge

terminated armchair BNNRs, however, remain semiconducting [174]. The chemically

Page 47: Graphene-analogous low-dimensional materials

46

modified armchair BNNRs by different transition metals can be semiconductors,

half-metals or metals with diverse magnetic ground states depending on the types of

metals [175]. These interesting band-gap modulations by edge decorations may be

exploited for nanoelectronic applications.

3.1.4 BN/graphene hybrid structures

Stimulated by the intensive studies of graphene and BN, the hybrid analogues of

graphene and h-BN, containing boron, nitrogen and carbon, have also attracted

intensive interests. Because of their commensurate structural parameters and distinct

electronic properties, graphene and BN are considered as good candidates for

fabricating B/N/C materials offering new functionalities. Other than forming solid

solution of B, N, and C, which has been studied extensively, it is necessary to explore

the possibility of making a graphene/BN composite, where the two phases coexist

separately, in a layered structure or in a same plane. Such a novel composite is our

main focus here.

A layered graphene/BN superlattice can be upheld together by vdW interactions,

such as placing graphene on single layer BN to form a bilayer system [176] and on

two or more BN layers to construct a multilayer heterostructure [177], or designing

3D superlattice with alternate stacking of graphene and BN monolayer [178]. Other

heterostructures such as monolayer [179] and bilayer [180] graphene sandwiched

between two BN layers or monolayer BN sandwiched between two graphene layers

[181] have also been constructed.

Page 48: Graphene-analogous low-dimensional materials

47

For graphene/BN bilayer, the most stable configuration has one C atom on top of

a B atom while the other C atom sites above the center of the BN hexagon ring, and a

small gap (53 meV) is opened around the K-point in graphene resulting from

symmetry-breaking of the two carbon sublattices [176] (Fig. 17). The band gap of

graphene/BN bilayer or multilayer structure can be modulated by tuning the interlayer

spacing and stacking pattern [182], hydrogenating [183], introducing strain [184], or

applying an electric field in the direction perpendicular to the hybrid layers [185].

Note that according to Xu et al.’s report [186], the response of electronic properties of

graphene/BN heterostructurs on the applied external pressure might lead to the

potential realization of atomic-scale pressure sensors.

Fig. 17. Model (a) and band structure (b) of single layer graphene on h-BN.

Reproduced from Ref. [176]. Copyright © 2007, American Physical Society.

Experimentally, many efforts were devoted to synthesize graphene/BN hybrid

heterolayers [187]. For example, Gannett et al. [188] and Kim et al. [189] separately

grew graphene on Cu foil through CVD technique, and then transferred the fabricated

graphene onto the exfoliated BN underlayer. Graphene has also been grown by CVD

Page 49: Graphene-analogous low-dimensional materials

48

onto BN monolayer on Ru(0001) [190] and Ni(111) [191] surfaces. Alternatively,

Bresnehan et al. [192] firstly grew large-scale BN layer on Cu foil by a catalytic

thermal CVD, and followed by the transfer of BN to quasi-freestanding epitaxial

graphene grown on SiC wafer. Besides the use of metal catalysts, Son et al. [193]

reported a direct catalyst-free growth of flat graphene pads by atmospheric CVD

process on top of mechanically exfoliated BN substrate. Lin et al. [194] introduced a

hydrogen flame synthesis method to successfully grow graphene on BN flakes using

PMMA as carbon sources. Very recently, Haigh et al. [195] have reported a

step-by-step preparation of graphene/BN multilayer heterostructures in which the

mono- and bilayer graphene were individually interlaid between atomically thin h-BN

crystals. As a representative case, Fig. 18a sketches the hybrid devices with two

graphene layers (dark grey) intercalated into thin BN layers (blue) deposited on top of

an oxidized silicon wafer (violet and grey). Fig. 18b shows the optical microscopy of

fabricated heterostructure, and the graphene Hall bar is highlighted with thin black

lines. The scanning electron microscopy (SEM) images of a cross-sectional specimen

extracted from the device are presented in Fig. 18c and d, and the red circles (known

as “bubbles”) indicate the trapped chemical species, mostly the surface-adsorbed

hydrocarbons. These adsorbates are detrimental to graphene performance, and can be

further eliminated by thermal annealing to obtain clean and atomically flat interface

[196]. The high-resolution scanning transmission electron microscopy (STEM) image

in Fig. 18e provides a useful tool to count the atomic layer sequences, and clearly, this

unique device is alternatively stacked with one graphene layer and four BN layers

Page 50: Graphene-analogous low-dimensional materials

49

(Fig. 18f).

(a) (b) (c)

(d)

(e) (f)

Fig. 18. (a) Schematic illustration that two graphene monolayers are interlaid with

h-BN crystals. (b) Optical image, (c) SEM micrograph, and (d) low-magnification

SEM image of the device. (e) High-resolution bright-field aberration-corrected STEM

image of the graphene-h-BN heterostructure. (f) Schematic of the atomic layer

sequence for this device. Reproduced from Ref. [195]. Copyright © 2012, Nature

Publishing Group.

Actually, because of the good lattice match and the atomically smooth surface

free of dangling bonds and charge traps, h-BN serves as an excellent back-gate

dielectric layer for graphene-based field-effect transistors [ 197 , 198 ], electron

tunneling devices [199], and is used as an appealing supporting-substrate to enhance

carrier mobilities of graphene electronics that have better quality than on SiO2 or

other high-k dielectrics substrates such as HfO2 and Al2O3 [200 - 202 ]. As a

comparison, Fig. 19 presents a side-by-side STM topography of typical graphene

corrugations on BN and SiO2 substrate [203], and they obviously reveal a reduced

roughness of graphene/BN over large areas than its graphene/SiO2 counterpart.

Moreover, the charge density inhomogenety in graphene/BN system is also

Page 51: Graphene-analogous low-dimensional materials

50

significantly reduced as compared with graphene/SiO2 system, which is clearly

observed from the charge density maps in Fig. 19c and d. Note that placing graphene

onto BN can actually yield improved device performance, yet contrary to the

theoretical prediction; the existence of gap-opening for graphene/BN system has not

been detected by spectroscopy measurements, which is mainly because of the stacking

misalignment of the graphene/BN lattices [204].

Fig. 19. Comparing topography and charge density for graphene/BN and

graphene/SiO2. Reproduced from Ref. [203]. Copyright © 2011, American Chemical

Society.

In another way, a small fraction of BN region can be substituted with C atoms or

vice versa to form hybridized single layer superlattice, and there has been some

success in preparing BxCyNz compositions. For example, based on CVD growth, Ci et

al. [76] fabricated hybrid BNC atomic sheets containing randomly and separately

distributed h-BN and C phases, as shown by the atomic model in Fig. 20. In their

experiment, methane and amine borane (NH3-BH3) were adopted as the precursors for

C and BN, respectively. By controlling the relative proportion of C and BN sources

Page 52: Graphene-analogous low-dimensional materials

51

during the CVD process, the atomic ratio of B, N, and C of BNC hybrids can be

flexibly tuned over a wide scale ranging from pure BN to pure graphene, and the

electrical properties can be accordingly controlled from insulators to conductors by

increasing the C concentration. Lin et al. [205] developed a method to convert

graphene oxide (GO) into BCN nanosheets through substitutional doping by reacting

GO with B2O3 and ammonia at 900-1100 C. Based on a sequential CVD growth,

Sutter et al. [206] prepared monolayer graphene-BN heterostructures on Ru(0001)

substrate by exposing the metal surface to high-purity ethylene and borazine,

respectively. Pakdel et al. [207] developed a non-catalytic CVD method for growing

boron nitride-carbon (BN-C) phase-separated composite nanosheet coatings on

Si/SiO2 substrates. A chemical blowing method which relies on making large bubbles

within thin B-N-C-H polymer walls was also used to produce Cx-BN nanosheets after

high-temperature annealing and collapse of the polymer bubbles [101]. Moreover,

Raidongia et al. [111] synthesized a graphene-like material with composition of BCN

by reacting high-surface-area activated charcoal with a mixture of boric acid and urea

at 900 C. This material mainly consists of two to three layers, showing a relatively

high surface area of 2911 m2g-1 and exhibiting a high propensity for CO2 adsorption

and hydrogen uptake. The BCN material exhibitd femtosecond third-order nonlinear

optical susceptibility and hot carrier dynamics, making BCN an attractive candidate

for untrafast optical applications [208].

Page 53: Graphene-analogous low-dimensional materials

52

Fig. 20. Atomic model of the h-BNC film showing hybridized h-BN and graphene

domains. Reproduced from Ref. [76]. Copyright © 2010, Nature Publishing Group.

Theoretically, hybrid graphene/BN single-layer sheets were predicted to display

tunable properties depending on the arrangement of the separated graphene or BN

regions and the compositional ratio of carbon to BN in the structure [ 218 ].

Structurally, graphene has a smaller interatomic distance (1.42 Å) than BN (1.45 Å),

and the incorporation of graphene into BN lattice would induce a tensile stress after

relaxation, which ensures the overall planarity of the C-BN composite system.

Because of quantum confinement, the intrinsically semiconducting graphene islands

embedded in the BN matrix gain the character of quantum dots [219]. Typically, a

hybrid BCN honeycomb sheet with triangle-shaped graphene quantum-dot shows

flat-band ferromagnetism, while the magnetism vanishes when the embedded

graphene adopts a hexagonal-shape quantum dot [220]. The magnetism mainly

originates from π electrons of the embedded graphene, especially located at the C

atoms along the zigzag edges of the graphene flakes. Essentially, the substitution of

one B atom by one C atom will bring one more electron into the system, while the

Page 54: Graphene-analogous low-dimensional materials

53

substitution of one N atom by one C atom will introduce one more hole, and the net

effect of the hybrid system becomes hole-injected or electron-injected if the

embedding border is connected with C-N (in N-rich condition) or C-B (in B-rich

condition) bonds. In both cases, the band gap of the hybrid architecture reduces with

increasing the embedded graphene size. On the other hand, the BN quantum dots

embedded into graphene lattices would lead to gap opening of graphene regardless of

sizes and geometries of the interior BN domains [221,222].

Structural hybridization by substituting graphene C-C pairs with isoelectronic

BN pairs also achieves significant tuning on the electronic transmission of 1D

graphene/BN nanoribbons [223,224]. Note that half metallicity can be realized in

graphene nanoribbons by applying an external electric field, yet a carefully designed

graphene-BN material such as zigzag graphene/BN nanoribbons with dihydrogenated

B edge and the zigzag graphene nanoroads embedded in BN can eliminate the need of

electric field to possess intrinsic half metallicity [225,226]. For BN-embedded zigzag

GNRs (Fig. 21) [227], a gradual replacement of the zigzag C-C chains in the middle

part by zigzag B-N chains transforms the system finally to the zigzag BN nanoribbons,

and the electronic structures vary accordingly with the doping concentration. At a

high doping concentration, the hybrid nanoribbons with terminated polyacene C

chains at the edge and all the substituted B-N chains in the internal part act as

half-metallic antiferromagnets for all the widths. Application of electric field on

zigzag BN nanoribbons with polyacene edges shows that the half-metallic behavior

sustains over sufficiently large electric field strength. The Lewis acid character of

Page 55: Graphene-analogous low-dimensional materials

54

boron is identified to be responsible for the charge transfer from adjacent C atoms to

the B atoms, resulting in an interface potential gradient analogous to the effect of

external electric field and invoking half metallicity. Furthermore, at a precise B-C-N

composition, the BC2N sheet is a direct-gap semiconductor with a band gap of 1.6 eV

which is intermediate between graphene and BN, and the derived BC2N nanoribbons

can be semiconducting, metallic, or half metallic, which is width- and edge-dependant

[228].

Fig. 21. A hydrogen passivated 8-ZGNR without (upper) and with (lower) BN doping.

Doping concentration increases gradually along the directions of solid arrows.

Electric field was applied along the y-axis direction. Reproduced from Ref. [227].

Copyright © 2009, American Physical Society.

Note that the C-BN hybridized phase not only shows great potential for band-gap

engineering, but also demonstrates promising applications in hydrogen storage since

the interface between graphene and BN exhibits high affinity for gas adsorption,

especially at the zigzag-type interface [111,229]. Beyond this, Wang et al. [230] have

recently showed that the BCN graphene exhibits efficient metal-free electrocatalytic

activity for oxygen reduction reaction (ORR) in fuel cells. Lei et al. [231] revealed

Page 56: Graphene-analogous low-dimensional materials

55

that the BCN nanosheets showed promising storage performance in lithium ion

batteries, and exhibited a stable capacity of ~100 mA h g-1 at 2 A g-1 for 5000 cycles.

3.1.5 Potential applications

The unique and exceptional properties with regard to the mechanical, electronic,

thermal, and chemical aspects favor BN nanosheets (BNNS) for use in a wide range

of electronic and composite applications. For example, h-BN nanosheets are widely

used as ultraviolet-light laser devices [81,232], field emitters [233], semiconductor

diodes [234], and insulating thermoconductive nanofillers in polymer or ceramic

composites [39,235]. For example, ultrathin BN nanosheets protruded from Si3N4

nanowires [233] or BN fibers [236] as well as porous BN nanospheres [237] display

excellent field emission performance with electron emission property comparable to

that of carbon nanotubes. BN layers also widely act as good insulating substrates for

graphene-based electronics, as discussed in the previous section.

On the other hand, due to its attractive combination of electrical insulation, high

thermal conductivity, and optical transparency, layered h-BN holds great promise in

applications as polymeric composite fillers. According to the experimental results, the

incorporation of exfoliated BN sheets into PMMA [39] and

poly[2,2'-(p-oxydiphenylene)-5,5'-bibenzimidazole (OPBI) [42] matrix showed

greatly improved mechanical and thermal properties as compared with the neat

polymers. As an example, Fig. 22 presented a comparative study on the thermal

mechanical analysis (TMA) and mechanical strength test of blank PMMA polymer

Page 57: Graphene-analogous low-dimensional materials

56

and its BN composite (with 0.3 wt% BN addition). By comparing with the parent

PMMA, the embedment of BN nanosheets in PMMA results in a remarkable

reduction of the coefficient of thermal expansion (CTE) and leads to enhanced elastic

modulus (22% increase) and strength (11% increase). Besides, Song et al. [238] have

recently showed that nanocomposites containing exfoliated BN nanosheets and

poly(vinyl alcohol) or epoxy resin polymers possess superior thermal transport

performance. Furthermore, the BN layers could be used as nanofillers in heat transfer

fluids. Taha-Tijerina et al. [239] have recently reported the synthesis of a stable

nanofluid with BN fillers in mineral oil (MO), a commonly used heat-transfer fluid in

transformers. This novel BN/MO nanofluid has a high thermal conductivity which

tends to increase with h-BN filler concentration, and only 0.1wt% addition of h-BN in

MO can achieve a significantly enhanced thermal conductivity of around 76% at 373

K.

Fig. 22. (a) Comparative TMA and (b,c) mechanical strength tests on blank PMMA

and its BN nanosheet composite. The inset in (a) shows CTE data before and after

Page 58: Graphene-analogous low-dimensional materials

57

glass-transition temperature. Reproduced from Ref. [39]. Copyright © 2009,

Wiley-VCH.

Moreover, functionalized BN nanosheets also demonstrated great potential to

fabricate polymer composites. In an experimental work, Liu et al. [240] recorded that

BN nanosheets functionalized by P(S-b-MMA) modifier serve as a better nanofiller to

improve the tensile strength of PMMA and polystyrene (PS) matrix than the

unmodified BN filler. Quiet recently, Sainsbury et al. [241] described a novel

functionalization strategy to yield hydroxyl-functionalized BN nanosheets. This

involves a two-step procedure: initially covalently grafting alkoxy groups to the B

atoms in h-BN lattice via oxygen radical attack, and the subsequent hydrolytic

defunctionalization of the alkoxy groups to generate surface OH-functionalized BN

nanosheets (OH-BNNSs) (Fig. 23a). Note that the bare BNNSs are typically

superhydrophobic, while surface modification by introducing hydrophilic OH

functional groups renders the nanosheets water-soluble. Fig. 23b presents photographs

of water, pristine BNNSs, and OH-BNNSs (from left to the right) under an irradiation

of a 532 nm green laser, which clearly shows the enhanced solubility and scattering

behavior of the OH-BNNSs relative to the pristine BNNSs. When used as polymer

nanofillers, the OH-BNNSs significantly reinforce the polyvinyl alcohol (PVA) in its

mechanical strength (Fig. 23c and d) over the pristine BNNSs, since the presence of

surface OH groups provide better chemical compatibilization and dispersability with

the host matrix.

Page 59: Graphene-analogous low-dimensional materials

58

Fig. 23. (a) Reaction scheme indicating the two-step procedure for oxygen radical

functionalization of BN nanosheets. (b) Photographs of water, pristine BNNSs and

OH-BNNSs in water. (c) Stress-strain curves for PVA reference and BNNS:PVA,

OH-BNNS:PVA composites. Inset shows the transparent samples on glass

microscope slides. (d) Young’s modulus, ultimate tensile strength, strain to break, and

toughness for the PVA reference and composites. Reproduced from Ref. [241].

Copyright © 2012, American Chemical Society.

In addition to the applications mentioned above, BN nanosheets are also

promising as an alternative for graphene to plane electronics where lithography

technique can be conveniently employed, as superhydrophobic surface coating, as

nanoscale support for metal and metal oxide catalysts. For example, the BN nanomesh

deposited onto transition metal surfaces is thermally very stable and serves as a

template for surface self-assembly of well-ordered organic molecules [242], water

molecules [243], rare gas atoms [244], transition metal atoms [245], or metallic

nanoparticles [246].

Page 60: Graphene-analogous low-dimensional materials

59

These substrate-supported molecular or atomic arrays can be used to produce

functional supramolecular BN nanostructures, which demonstrate great potential in

catalytic and sensing applications. For instance, a theoretical study by Gao et al. [247]

suggested that Au supported on BN surface exhibits enhanced adsorption and

catalytic activation for O2 molecule. Experimentally, Wang et al. [248] showed that

Au- and Pt-nanoparticles loaded onto BN nanosheets function as efficient catalysts

towards CO oxidation. Wang et al. [249] also showed that Au nanoparticle-loaded

h-BN nanocrystals can be used for oxidation of benzyl alcohol to benzaldehyde. Apart

from the catalytic applications, Lin et al. [250] reported the fabrication of Ag

nanoparticle-decorated BN nanosheets, which were further transferred onto quartz

substrates to fabricate reusable and oxidation-resistent surface enhanced Raman

spectroscopy (SERS) sensor devices.

Note that the 1D BN nanoribbons are theoretically predicted to have significant

potentials in the area of nano-electronics and spintronics; however, their realistic

applications remain currently less explored by experimentalists. This challenge is

possibly due to the large experimental difficulty in producing high-quality BNNRs

with uniform and smooth edges.

3.2 Silicene

Silicene, a one-atom-thick silicon sheet arranged in a 2D honeycomb lattice, is a

new allotrope of silicon, similar to graphene. Due to their structural resemblance,

single-layer silicene is also alternatively considered as a Si-based 2D counterpart of

Page 61: Graphene-analogous low-dimensional materials

60

graphene. However, unlike graphene, silicene favors to adopt a low-corrugated

honeycomb configuration (about 0.44 Å buckling is present) [251] rather than a planar

one, forming a mixed sp2-sp3 like hybridized state [252]. This difference stems from

the fact that for C atoms, the fully sp2-hybridized state is more stable than the

sp3-hybridized state, while the case for Si is reverse. This could also be otherwise

understood by the instability of Si atoms to form strong Si-Si π bonds to stabilize the

planar three-fold coordination. Due to its unique bonding behavior, bulk Si can not

form a layered phase like graphite, yet one exception is that the layered Si atoms can

survive in alkaline-earth-metal silicides (such as CaSi2), which are structurally

composed of alternative layer-by-layer packing of hexagonal alkaline-earth-metal

layers and corrugated Si(111) atomic layers.

3.2.1 Synthesis of silicene and functionalized silicene

The possible existence of silicene has been theoretically conjectured since 1994

and has been just synthesized in recent years [253]. Note that silicene is a meta-stable

structure, which is 1.56 eV/atom less stable than bulk Si. This indicates that growth of

silicene necessitates the use of deposition substrates. Actually, although free-standing

silicene has not been realized up to now, experimental evidence for the existence of

epitaxial silicene does exist, which has been discussed in detail in the previous

Section 2.4.

At present, the silicene structures are obtained mainly by the surface-assisted

epitaxial growth, such as the observed silicene nanoribbons on Ag(110) [84-86] and

Page 62: Graphene-analogous low-dimensional materials

61

silicene sheets on Ag(111) [88-93], Ir (111) [94], or ZrB2(0001) surface [95]. Unlike

the high oxidation reactivity of typically sp3-bonded Si surface, De Padova et al. [254]

discovered that the 1D grafting of silicene nanoribbons grown on Ag(110) surface are

room-temperature oxidation-resistant, which compares favorably with graphene.

Dávila et al. [255] have recently showed that isolated or self-aligned Si nanoribbons

on Ag(110) can be easily attacked by atomic or molecular hydrogen adsorption.

Note that even prior to the synthesis of silicene, 2D functionalized Si nanosheets

have been prepared by Nakano et al. using solution-phase methods. For example,

layered siloxene [Si6H3(OH)3], which is composed of corrugated Si layers with Si

atoms being terminated by H and OH groups, alternatively, was exfoliated by using

sodium dodecyl sulfate as dispersant [256]. To prepare oxygen-free modified Si

nanosheets, they performed the exfoliation of layered polysilane (Si6H6) as a result of

reaction with n-decylamine, and the resulting Si nanosheets are covalently covered

with organic amine groups [257]. In particular, they also synthesized oxygen-free,

phenyl-modified organosilicon nanosheet (Si6H4Ph2) by reacting the layered Si6H6

with PhMgBr [258] (Fig. 24). The yielding Si6H4Ph2 exhibits a completely flat plane

surface, with a measured sheet thickness of 1.11 nm and a periodic distance of 0.94

(0.98) nm of surface phenyl groups, in consistency with the thickness (0.98 nm) and

periodicity (1.0 nm) calculated on the basis of its atomic architecture. Very recently,

Nakano et al. have also synthesized alkyl-modified Si nanosheets through

hydrosilylation of layered Si6H6 [259]. Structurally, these modified Si nanosheets are

the functionalized derivatives of 2D silicene structure, which can be regarded as a

Page 63: Graphene-analogous low-dimensional materials

62

new class of 2D functional Si materials. Note that unlike the bare silicene sheet, the

functionalized Si nanosheets can stably exist as a free-standing form, and become

soluble in organic solvents. Thus they are likely to be deposited onto various

substrates for characterization and application.

Fig. 24. (a) AFM image of [Si6H4Ph2]. (b) Line profile along the black line in (a). (c)

Model of [Si6H4Ph2]. (d) AFM image of the surface of [Si6H4Ph2]. (e) Line profile

along the black line in (d). (f) Top view of the model structure for [Si6H4Ph2].

Reproduced from Ref. [258]. Copyright © 2010, American Chemical Society.

3.2.2 Theoretical investigations of silicene

The free-standing silicene sheet and its nanoribbons have been widely

investigated through first-principles computations, and their electronic properties bear

great resemblance to those of graphene.

2D silicene. Structually, the buckled Si hexagonal sheet is optimized to have a

lattice constant of a = 3.86 Å, and the nearest Si-Si distance is relaxed to be 2.28 Å.

Page 64: Graphene-analogous low-dimensional materials

63

Compared with the C-C distance (1.42 Å) in graphene, such large Si-Si distance leads

to greatly weakened π-π overlaping, and thus the planar Si structure can not be

stabilized. Due to its structural corrugation, silicene has a lower D3h symmetry instead

of the D6h symmetry of graphene (Fig. 25a). Despite this, they share essentially

similar electronic structures and consequently analogous physico-chemical properties.

Electrically, silicene is a gapless semimetal, and its charge carrier behaves like a

massless Dirac fermion due to the crossing of π and π* bands at the Fermi level,

similar to graphene (Fig. 25b). Note that although the free-standing silicene is

expected to have a zero gap, a minigap can be opened in epitaxial silicene, resulting

from the symmetry-breaking induced by the interaction with the Ag substrate [260].

The existence of Dirac ferumions for silicene on Ag(111) has been observed by

angular-resolved photoelectron spectroscopy (ARPES) [89] and scanning tunneling

spectroscopy (STS) [261], and an Fermi velocity of vF = 1.3 × 106 ms−1 is obtained,

which is comparable to that found for graphene. Besides, the graphene-like Dirac

cones have also been observed by ARPES in monolayer and multilayer silicene

nanoribbons grown on Ag(110) [85,86] .

Fig. 25. (a) Top and side view of silicene. (b-d) Band structures of silicene around Ef

at three different vertical electric fields. Reproduced from Ref. [263]. Copyright ©

2011, American Chemical Society.

Page 65: Graphene-analogous low-dimensional materials

64

Because of their similarity in electronic properties, all graphene expectations,

such as high-speed nanoelectronic devices based on ballistic transport, can be

consequently transferred to silicene due to its high compatibility with the current

Si-based semiconductor industries. Moreover, since Si atoms have greater intrinsic

spin-orbit coupling strength than C atoms, silicene with topologically nontrivial

electronic structure is theoretically predicted to have a spin-orbit band gap of 1.55

meV that is much larger than that of graphene, and exhibit quantum spin Hall effect in

an accessible temperature regime [262]. This property makes silicene particularly

interesting for applications as spin Hall effect devices.

Silicene is a versatile material, because not only it owns the fascinating

properties of graphene, but also the structural puckering leads to a more reactive

surface, making band-gap-tuning of silicene much easier, which will be discussed in

the following part.

The electronic structures of silicene can be tuned by external electric field. For

example, Ni et al. [263] revealed that the band gap opens in buckled silicene by

applying a vertical electric field, and the size of the band gap increases linearly with

electric field strength. Specifically, the opened band gap for silicene is 0.08 eV under

E = 0.51 V/Å and becomes doubled under E = 1.03 V/Å (at the GGA/DNP level) (Fig.

25c and d). It is also remarkable that a topological quantum phase transition is

induced in silicene by changing the electric field [264]. The main effect of electric

field is to break the symmetry in A and B sublattices of silicene’s honeycomd and

hence leads to gap opening in the band structure. It should be recalled that by

Page 66: Graphene-analogous low-dimensional materials

65

imposing an external electric field graphene still remains its zero-gap semimetallic

character since its two sublattices remain equivalent. As a result, the biased

monolayer graphene did not function effectively as a field effect transistor (FET). For

comparison, the effective band-gap response of monolayer silicene as a function of

external electric field makes it a more suitable candidate for FET devices.

In addition, silicene can be alternatively modified by surface functionalization.

For example, hydrogenation opens a band gap in silicene, and the fully hydrogenated

silicene (also known as polysilane or silicane) with a composition of SiH is a

wide-band-gap semiconductor, with its band gap ranging between 2 (DFT-LDA level)

and 4 eV (GW approximation) [265]. Similar to graphene, the fully hydrogenated

silicene also prefers the chair-like configuration, but exhibits an indirect band gap.

Interestingly, surface decoration by halogen atoms (F, Cl, or Br) changes the indirect

gap of chair-like polysilane into the direct one, and the resulting gap value (1~2 eV) is

predicted to be smaller than the hydrogenated counterpart [266]. For the partially

hydrogenated silicene, ferromagnetism is introduced by attaching hydrogen atoms on

one side of the sheet (semihydrogenated silicene) [267]. The semihydrogenated

silicene is identified as a magnetic semiconductor with an estimated gap of 0.94 eV

(DFT-GGA level, Fig. 26), and is corrected to be 1.74 eV with HSE06 functional.

Osborn et al. [ 268 ] also studied the partial hydrogenation of silicene with

hydrogenation ratio ranging between 3.1% and 100%, and showed that by specifically

patterning the silicene with H domains, a metal-semiconductor-insulator functionality

can be obtained.

Page 67: Graphene-analogous low-dimensional materials

66

Fig. 26. Top (a) and side (b) view of semihydrogenated silicene, and its band structure

(c). Reproduced from Ref. [267]. Copyright © 2012, American Chemical Society.

Moreover, surface adsorption of metal atoms provides a variety of ways to

functionalize and modify the electronic properties of silicene. Lin et al. [269] revealed

that metal atoms, such as alkali and alkali-earth metals (Li, Na, K, or Ca), groups III

and IV metals (Al, Ga, In, or Sn), and transition metals (Ti, Fe, Co, Ni, Pd, Pt, or Au),

form strong interactions with silicene, and the formed complexes exhibit rich

electronic properties with the introduced s or d electron states lying above or below

the Fermi level. The strong binding of metal adatoms to silicene makes it potentially

useful in gas storage, superconductivity and catalysis. Osborn et al. [270] additionally

studied the Li chemisorption on silicene, and showed that the completely lithiated

silicene (composition of SiLi) is a stable and semiconducting material with a band gap

of 0.37 eV.

1D silicene nanoribbons. Like graphene, the applications of silicene in future

nanoelectronics would need a band gap. In our above discussion, the electric field

engineering only provides an external control, and the surface functionalization would

induce the change of hybridized state into sp3. Another effective method to achieve

Page 68: Graphene-analogous low-dimensional materials

67

gap opening in silicene is to reduce the dimensionality into 1D nanoribbons.

Due to the presence of quantum and edge effects, silicene nanoribbons exhibit

intriguing properties distinct from those of silicene, which are electrically width and

edge dependent [251]. Because of the instability of edge dangling bonds, the bare

armchair or zigzag edges undergo a sharp (2×1) reconstruction (Fig. 27), while the

reconstruction disappears when the edges are terminated by H atoms. The armchair Si

nanoribbons, with either bare or H-terminated edges, are all non-magnetic

semiconductors. Similar to armchair GNRs, their semiconducting gaps exhibit an

oscillatory behavior, in which the band gap is smaller for n = 3p + 2 than for n = 3p or

3p +1 (n is an integer) (Fig. 27 b and c). Zigzag Si nanoribbons, on the other hand, are

metallic or semiconducting depending on whether their edges are saturated or not.

Particularly, the bare zigzag ribbons have a non-magnetic and metallic ground state

(Fig. 27d). After H-saturation, the antiferromagnetic (opposite spin on the two

different edges) semiconducting state becomes most stable (Fig. 27e), and the band

gap of zigzag Si ribbons decreases monotonically as the width increases (Fig. 27f),

sharing great similarity with zigzag GNRs [271]. Like the fully hydrogenated GNRs,

the fully hydrogenated armchair- or zigzag-silicene nanoribbons are all non-magnetic

semiconductors [272].

The electromagnetic behaviors of zigzag Si nanoribbons can be modified by

external electric field or magnetic field. Especially, when a transverse electric field is

applied, the zigzag Si ribbons become half-metals [271], and a spin-polarized current

can be generated [273]. By applying a vertical magnetic field, Xu et al. [274] found

Page 69: Graphene-analogous low-dimensional materials

68

that zigzag Si nanoribbons can switch between anti-FM and FM coupled states, and a

large magnetoresistance (MR) is obtained due to the large current difference between

the semiconducting anti-FM state and the metallic FM state. This suggests potential

technological applications of silicene in spin-valve devices [275].

Fig. 27. (a) Ideal and relaxed bare 10-armchair Si nanoribbon. Band gaps as a

function of width for bare (b) and H-terminated (c) Si armchair nanoribbons. (d) Bare

Si zigzag nanoribbons showing two different (2×1) reconstruction indicated by ‘‘1’’

and ‘‘2’’ and the band structure corresponding to ‘‘1’’. Reproduced from Ref. [251].

Copyright © 2009, American Physical Society. (e) Spin density distribution of

H-terminated 6-zigzag Si nanoribbon at anti-FM state and its band structure. (f) Band

gaps of H-terminated zigzag ribbons vs the number Nz of ribbon width. Reproduced

from Ref. [271]. Copyright © 2009, American Institute of Physics.

Page 70: Graphene-analogous low-dimensional materials

69

3.2.3 Potential applications of silicene

As an important structural analogue of graphene, silicene possesses extraordinary

electronic properties, and its reactive surface can be modified by many methods.

Besides, it is advantageous that silicene can be easily integrated into Si-based

semiconductor industries. Although the current progress on silicene is still at its

infancy, silicene is expected to be developed for many technical prospects [276]. For

example, due to its high electron mobility, an intrinsic opening of spin-orbital band

gap, and tunable electronic structures, silicene can be used in nanoelectronics,

spintronics, and photonic devices. Thermal transport computations by Pan et al. [277]

suggested that armchair silicene nanoribbons may be used as high performance

thermoelectric materials. Hu et al. [278] showed that silicene with divacancy defects

exhibits high selectivy for H2 over H2O, N2, CO, CO2, and CH4, and can be

potentially used as hydrogen purification membrane. Moreover, since Si

nanomaterials have been largely explored for Li-ion battery anodes due to their high

energy density, silicene sheets should also be a promising anode candidate for Li-ion

batteries, because of its large specific surface area allowing for occupation of more Li

ions and contributing to a higher energy density.

Note that unlike graphene which can form a free-standing sheet, silicene has only

been found to grow on metal substrates. However, to promote silicene into the next

pace for nanoelectronic infrastructure, it is of great importance to grow silicene on

insulator substrates, and these synthesis challenges would encourage more theoretical

and experimental studies in the future.

Page 71: Graphene-analogous low-dimensional materials

70

3.3 BC3 honeycomb sheets

BC3 is a new graphite-like material with a layered hexagonal structure, which

can be viewed as the substitutional boron doping of graphite. Bulk BC3 was

successfully synthesized by reacting benzene and boron trichloride at about 800°C

[279]. Recently, the uniform BC3 sheets have been prepared in an epitaxial way onto

NbB2(0001) substrate via C substitution in the B honeycomb layer of metal diboride

[280], and the phonon dispersion curves of BC3 sheets have been determined [281].

Compared with pure carbon materials, BC3 has the promising technological benefits

such as high oxidation resistance and superconductivity [282].

Fig. 28. Atomic arrangement of BC3. Reproduced from Ref. [283]. Copyright © 2012,

American Chemical Society.

The most stable ordered structure at the BC3 composition is composed of carbon

hexagons and orderly distributed boron atoms, where six carbons construct a hexagon

and each boron atom is linked to three separated hexagons. Although boron is

electron-deficient as compared with carbon, the charge in BC3 sheet is actually

transferred from boron to carbon. Note that in contrast with graphene where the π

electrons are delocalized over entire carbon lattice, the BC3 lattice displays the

Page 72: Graphene-analogous low-dimensional materials

71

aromatic carbon rings with six π-electrons located over each carbon hexagon, and are

separated by the anti-aromatic hexagons without π-electrons consisting of carbon and

boron atoms (Fig. 28) [283].

A number of studies have examined the electronic properties, defects, doping,

and surface functionalization of BC3 sheet. BC3 is a non-magnetic semiconductor with

an indirect LDA band gap of 0.54 eV. By inducing defects (vacancy, or antisites), the

electronic and magnetic properties of BC3 sheet can be tailored remarkably [284]. Lin

et al. [285] showed that by substituting the B or C atoms in BC3 lattice with transition

metals (Fe, Co, or Ni), a rich variety of properties ranging from magnetic

semiconductor, nonmagnetic semiconductor, to magnetic metal can emerge. Note that

generation of O-defects in BC3 sheet has been detected by angle-resolved x-ray

absorption near edge structure (XANES) after the sample is subject to Ar+ ion

bombardment and subsequent exposure to air [286]. The O-defects are formed by

firstly creating C vacancies and then substituting the vacancy sites with O atoms.

Moreover, it is possible to tailor electronic structures of BC3 by hydrogenation.

Through step-by-step hydrogenation, a semiconductor→metal transition appears in

hydrogenated BC3 sheet [ 287 ]. Ding et al. [ 288 ] further demonstrated that

half-metallicity is achieved in semihydrogenated BC3.

BC3 nanoribbons also show a rich variety of electronic properties. According to

Ding et al.’s report [289], armchair BC3 nanoribbons are all semiconductors, while

zigzag nanoribbons can be semiconductors or metals depending on the edge

geometries. Specifically, the zigzag nanoribbons with BCBC terminations are

Page 73: Graphene-analogous low-dimensional materials

72

semiconductors, and those with BCCC or CCCC edge terminations are metallic.

Additionally, Dutta et al. [290] showed that for the case of armchair BC3 nanoribbons,

removal of the passivating hydrogen atoms from the edge B atoms leads to a higher

stability and makes the narrow ribbons become metallic, and then a transition to

semiconductor occurs as the ribbon width increases. Because of the rich electronic

properties, BC3 and its derivatives are potential low-dimensional materials in

nanoelectronics.

The intriguing BC3 material demonstrates promising applications in hydrogen

storage and Li-ion batteries. For example, Zhang and Alavi [291] as well as Sha et al.

[292] predicted that bulk BC3 can be used as potential hydrogen storage candidate.

Compared with the large energy barrier for H2 diffusion into crystallite graphite, bulk

BC3 exhibits higher reactivity towards H2 and leads to reduction of the activation

energy for H2 migration. Within the BC3 lattice, the intercalated H2 molecule

undergoes dissociative chemisorption to form C-H bonds, and the adsorption strength

and H migration barrier are modest, making layered BC3 suitable for reversible

hydrogen storage under near-ambient conditions. Except for bulk BC3, Yang et al.

[293,294] showed that metal atoms (Li, or Ca) can be strongly adsorbed on BC3

monolayer without clustering, and Li-doped and Ca-doped BC3 sheets are the

favorable candidates for hydrogen storage. Besides, BC3 can also be promising as an

anode material for Li-ion batteries. Based on Kuzubov et al.’s computations [295], the

layered BC3 sheets are predicted to be the Li+ intercalation hosts with the Li ion

capacity (corresponding to Li2BC3 stoichiometry) surpassing that of graphite.

Page 74: Graphene-analogous low-dimensional materials

73

4. Hypothetical planar graphene analogues

4.1 SiC Silagraphene

For Si-C phases, a typical material with considerable interests is SiC. Si and C in

bulk SiC favor sp3 hybridization, and SiC naturally occurs in the form of cubic,

hexagonal and rhombohedral structures. Recently, a planar phase of SiC with

sp2-hybridized feature resembling graphene was theoretically predicted to have high

structural stability, and the electronic properties of this graphene-like material have

been studied [27,296].

The graphene-like SiC sheet consists of alternating Si and C atoms with each Si

atom having three C atoms as its nearest neighbors and vice versa, and the Si-C bond

length is optimized to be 1.79 Å. Due to the Si-C ionicity, monolayer SiC is a

semiconductor with a direct gap of about 2.55 eV (Fig. 29a), which is enlarged to be

3.7 eV after GW correction [297]. The SiC nanoribbons with H-passivated edges

possess interesting behaviors. The armchair SiC nanoribbons are non-magnetic

semiconductors, and the resulting direct band gaps exhibit sawtooth-like oscillation

features (Fig. 29c). The zigzag SiC nanoribbons are magnetic metals except for

thinner ribbons with small direct band gaps [27,296]. The spin-polarized magnetism

comes from the edge Si and C atoms (Fig. 29e). However, the two zigzag edges are

not antiferromagnetically coupled as in zigzag GNRs, but are ferrimagnetically

coupled, where the ferromagnetic chains at the two edges have opposite orientations

of magnetic moments but of different values, and thus the net magnetic moments for

the zigzag systems are not zero. This is ascribed to the unequal values of local

Page 75: Graphene-analogous low-dimensional materials

74

magnetic moments at the edge C and Si atoms. Interestingly, Sun et al. [27] reported

that zigzag SiC nanoribbons narrower than 4 nm present half-metallic behaviors (Fig.

29d). Based on HSE functional computations, Ding et al. [298] have recently revealed

that asymmetric H-terminations can break the magnetic degeneracy in zigzag SiC

nanoribbons and favor the ferromagnetic state, and meantime, the nanoribbons with

bare Si edges or two-hydrogen saturated Si (C) edges are converted into half metals.

Fig. 29. (a) Geometry and band structure of SiC sheet. (b) Structures of armchair and

zigzag SiC nanoribbons with width W. (c) Band gap of armchair SiC nanoribbons as a

function of ribbon width W. (d) Band structure of zigzag SiC nanoribbons with W=7.

Dotted red lines are spin-up bands, and solid blue lines are spin-down bands. The

projected band structure of 2D SiC sheet is shown by shaded areas. (e) Spin densities

for the W=7 zigzag SiC nanoribbon. Reproduced from Ref. [27]. Copyright © 2008,

American Institute of Physics.

The electronic structures and magnetization of SiC sheets and nanoribbons can

Page 76: Graphene-analogous low-dimensional materials

75

be further manipulated by other strategies such as transverse electric field [299],

surface modification [300], strains [301], defects [302], and doping [303,304]. For

example, Lou et al. [299] found that the narrow zigzag SiC nanoribbons are turned

into ferromagnetic metals from ferrimagnetic semiconductors when an external

transverse electric field is applied. Substitutional impurities, adatom adsorption and

vacancy defects have been demonstrated to yield controllable functionalization of SiC

honeycomb structure [302]. Moreover, carrier doping, such as hole, electron, or

chemical doping (B or N doping), is used to manipulate magnetism of narrow zigzag

SiC nanoribbons [303,304]. Like graphene, BN or silicene, hydrogen is also an

important surface modifier for SiC. The fully hydrogenated SiC sheet in its stable

chair-like structure is a non-magnetic semiconductor with a direct band gap of 3.84

eV (DFT-GGA) [305], enlarged to be 5.3 eV at the GW approximation [297]. For the

semihydrogenated case, Xu et al. [300] reported the induced FM and anti-FM

properties in SiC when C and Si atoms are hydrogenated, respectively. Surface

hydrogenation also plays a crucial role in tuning the electronic properties of SiC

nanoribbons. Guan et al. [ 306 ] reported that by carefully controlling the

hydrogenation ratio, a transition of the anti-FM semiconductor→FM metal→anti-FM

half metal→NM semiconductor is achieved in zigzag SiC nanoribbons.

Here SiC only represents one example of the graphene-like binary compounds of

group-IV elements. Many other binary phases, such as GeC, SnGe, SiGe, SnSi, and

SnC, have been theoretically studied on their structural, electronic and mechanical

properties [307]. Among their optimized monolayer structures, GeC and SnC form a

Page 77: Graphene-analogous low-dimensional materials

76

planar honeycomb structure like SiC, whereas SnGe, SiGe, and SnSi prefer a

low-buckled geometry for stabilization like silicene, with the buckled distance in the

range of 0.55~0.73 Å. Electrically, all these binary phases are predicted to be

semiconductors. Depending on the structural composition, their LDA band gaps vary

over a wide scale from 0.02 eV (SiGe) to 2.09 eV (GeC) [307].

4.2 SiC2 Silagraphene

The Si atoms in the above planar SiC sheet are three-fold coordinated. Motivated

by the intensive studies on planar tetracoordinate C (ptC) (known as the anti-van't

Hoff/Lebel compound), planar tetracoordinate Si (ptSi) chemistry has also invoked

tremendous interests for scientists [308]. From a theoretical perspective, Li et al. [29]

have recently predicted an interesting phase of SiC2 silagraphene where ptSi can be

stably embedded into periodic 2D graphitic network.

SiC2 sheet is designed based on a planar SiC4 molecule containing a ptSi (Fig.

30a). SiC4 has C2v symmetry, and is a local minimum at the potential energy surface.

Using planar SiC4 as the building unit, 2D infinite SiC2 sheet in an ideally planar

arrangement is accordingly constructed (Fig. 30b). In SiC2 silagraphene, each Si atom

is bonded with four C atoms to form a ptSi moiety, while each C atom is bonded with

two ptSi atoms and one C atom. The Si-C bond (1.916 Å) is slightly longer than the

typical Si-C single bond (1.87-1.91 Å), while the C-C bond (1.332 Å) is characterized

as a typical C=C double bond. Structurally, one C=C bond and its four neighboring Si

atoms resembles the structure of ethylene. In a reasonable sense, the

Page 78: Graphene-analogous low-dimensional materials

77

planarity-preferred ethylene-like skeletons act as a driving force to yield a perfectly

planar SiC2 network. The SiC2 silagraphene is a stable phase, as a local minimum at

the potential energy surface, having a large binding energy (6.04 eV/atom) which is

almost in coincidence with that of the SiC silagraphene, and possessing high kinetic

(up to 800 K) and thermodynamic stabilities.

Fig. 30. (a) B3LYP/6-311+ G* optimized structure of SiC4 molecule; the distances

are in angstroms. The number in bracket is computed from MP2 procedure. (b) Top

and side views of SiC2 silagraphene. (c) Band structure and partial DOS of SiC2

silagraphene. Reproduced from Ref. [29]. Copyright © 2010, American Chemical

Society.

The electronic properties of SiC2 related nanomaterials are in stark contrast to

their carbon and SiC analogues. All the ptSi-containing nanomaterials, ranging from

SiC2 silagraphene sheet to nanoribbons, are robustly metallic. Analyzing partial DOS

of SiC2 silagraphene reveals that the states at the Fermi level are mainly contributed

by C-2p and Si-3p states, and the contribution of C-2p is more pronounced than Si-3p,

thus the characteristic metallicity of SiC2 is governed by the π-type C=C bonds (Fig.

Page 79: Graphene-analogous low-dimensional materials

78

30c).

4.3 Boron sheets

Boron, having one electron fewer than carbon, has always been fascinating the

research community due to its richness of bonding chemistry. The natural crystalline

structure of boron is the α-rhombohedral solid, where the boron atoms form

symmetric 12-atom icosahedral clusters in a crystalline state. Though boron has only

three valence electrons, it can still undergo sp2 hybridization like carbon. The

existence of planar or quasiplanar boron clusters in the 10- to 15-atom range has been

evidenced in recent experiments [309]. However, no evidence of 2D planar structure

exists in its crystals since they are mainly built from B12 icosahedra, and as a

consequence, searching for a planar boron sheet with high structural stability remains

an intriguing but challenging task.

As a matter of fact, researchers have theoretically proposed several models of 2D

boron sheets. In contrast to carbon, the graphene-like hexagonal B monolayer is

unstable and tends to be distorted. This is mainly due to the fact that the

electron-deficient characters of boron tend to promote boron atoms to interact with

each other in forming multicenter bonds (usually forming three-center bonds with

B-B-B structural units). Herein, it shoud be mentioned that although the free-standing

graphene-like boron sheet is unlikely to exist, boron can actually form hexagonal

layered structures in AB2-type materials (such as MgB2, AlB2, BeB2, and TiB2), in

which boron atoms carry a negative charge of 1 and thus acquire configuration

Page 80: Graphene-analogous low-dimensional materials

79

similar to that of carbon. Similarly, theoretical studies made by Zhang et al. [310]

indicated that the hexagonal geometries can be stabilized on metal surfaces (Mg, Al,

Ti, Au, and Ag). Considering the instability of the standalone hexagonal phase, Evans

et al. [311], Cabria et al. [312], and Kunstmann et al. [313] later independently

proposed that boron sheet constructed by a planar triangular lattice with B7 building

units (filled hexagons with interior boron atoms at the center) is more stable than the

hexagonal boron sheet. However, the six-fold coordination of boron atoms in the

triangular phase is not compatible with the p-orbital symmetries, and the flat plane is

also energetically unstable with respect to buckling which breaks the triangular

symmetry (constructed as hexagonal pyramidal B7 units). Hence in the cases of both

hexagonal and triangular sheets, distortions provide the stability of 2D boron sheet.

Later, after Szwacki et al. [314] reported an unusually stable and highly

symmetric B80 fullerene in 2007, Tang et al. [315] immediately proposed a novel

boron monolayer comprised of a hybrid of triangular and hexagonal motifs, known as

α sheet. The α-sheet has been predicted to be energetically more stable than previous

triangular or hexagonal boron sheet (Fig. 31) [316]. This sheet is ground-state planar

and preserves the symmetry of the triangular lattice, and the binding energy is ~93%

of that of bulk α-rhombohedral solid. In α boron sheet, about 2/3 of the hexagons are

occupied by additional boron atoms, whereby 3/4 of the boron atoms have five nearest

neighbors and 1/4 of the boron atoms have six nearest neighbors, and the hollow

hexagons are evenly and symmetrically distributed within the lattices. The stability of

α-sheet is explained through balance between two-center bonding in hexagonal

Page 81: Graphene-analogous low-dimensional materials

80

regions and three-center bonding in triangular regions. Specifically, the hexagonal

sheet is an electron-deficient system with part of the in-plane sp2-bonding states

unoccupied, thus acts as an electron-acceptor. While the triangular sheet has a surplus

of electrons in anti-bonding states, thus acts as an electron-donor. After forming the

mixed hexagon-triangle lattice, the electrons provided by the triangular regions

occupy those empty bonding states, thereby markedly enhancing the structural

stability.

Fig. 31. Models of various 2D B sheets. The stability order is: α-sheet > β-sheet >

buckled triangular sheet > perfect triangular sheet > distorted hexagonal sheet >

perfect hexagonal sheet. Reproduced from Ref. [316]. Copyright © 2008, American

Chemical Society.

As further additions to α-sheet, other B sheets based on hybrid hexagonal and

triangular motifs have been proposed (Fig. 32), and some of them are comparable or

even more stable than the known α-sheet. For example, Tang et al. [317] revealed that

Page 82: Graphene-analogous low-dimensional materials

81

for single-layered B sheets made of hexagons and triangles, their ground-state

configurations may be either buckled or flat depending on the hexagon-to-triangle

ratio, or the “hexagon hole density” η. They showed that flat B sheets are located at

the range of 1/9 < η < 1/5, while otherwise the sheets stay buckled. Note that the

α-sheet has a η value of 1/9. According to Lau et al.’s opinion [316], the α-sheet is

more stable than the planar β-sheet (Fig. 31), which has η value of 1/8. Zope et al.

[318] designed a planar snub B sheet with a hexagon hole density of 1/7, and the

triangular regions in snub sheet form extended zigzag stripes. The proposed snub B

sheet is about 0.02 eV/atom less stable than the α-sheet. Besides, Özdoğan et al. [319]

proposed a γ-sheet, which is decorated similarly to the α-sheet but with parallel

hexagonal holes, yet the α-sheet is still comparatively more stable. Furthermore,

Penev et al. [320] discovered another two ground-state structures of planar B layers

that are as stable as the α-sheet but have higher hexagon density (1/8 and 2/15). More

recently, Yu et al. [321] have predicted a meta-stable B sheet (struc-1/8 B sheet),

which is 0.01 eV/atom less stable than α-sheet or the 1/8- and 2/15-B sheet.

Especially, Wu et al. [322] found that compared with the originally planar α-sheet, the

slightly buckled α-sheet (named α'-sheet) becomes energetically more stable;

moreover, their search yields two highly stable B monolayers, the planar α1-sheet and

β1-sheet, respectively, both with a hole density of η = 1/8. The α1-sheet and β1-sheet

are predicted to possess greater cohesive energies (by more than 60 meV/atom) than

the α-sheet or the 1/8- and 2/15-B layers. These proposed B monolayers with

comparable stabilities (particularly α-sheet, 1/8-sheet, 2/15-sheet, buckled α'-sheet,

Page 83: Graphene-analogous low-dimensional materials

82

α1-sheet, and β1-sheet) might constitute the leading candidates for the low-energy

configurations of 2D B sheets.

Fig. 32. Various structures of B monolayers based on hybrid hexagon-triangle motifs:

snub B sheet (Reproduced from Ref. [318]. Copyright © 2010, Elsevier.), γ-B sheet

(Reproduced from Ref. [319]. Copyright © 2010, American Chemical Society), 1/8-

or 2/15-B sheet (Reproduced from Ref. [320]. Copyright © 2012, American Chemical

Society.), struc-1/8 B sheet (Reproduced from Ref. [321]. Copyright © 2012,

American Chemical Society.), as well as α , α1, and β1-B sheet (Reproduced from Ref.

[322]. Copyright © 2012, American Chemical Society.). The red and yellow balls in

α -B sheet denote boron atoms moving outward or inward from the plane.

In terms of the electronic structure, nearly all the 2D boron monolayers are

rigorously metallic, with the exception of α- and α'-sheets. It is noteworthy that

Page 84: Graphene-analogous low-dimensional materials

83

although α- and α'-sheets are predicted to be metallic based on PBE functionals, the

hybrid PBE0 functionals suggest that both of them are semiconducting with an

indirect gap of 1.4 and 1.1 eV, respectively [322].

The electronic properties of α-B nanoribbons and potential applications of α-B

sheet have been studied. The α-B nanoribbons with line or zigzag edges are

electrically metallic; however, the presence of semiconducting behaviors is observed

in the two-hydrogen saturated armchair and zigzag boron nanoribbons [ 323 ].

Moreover, due to the unique geometry, the hexagonal regions of the α-B sheet can

interact strongly with alkali metal atoms (Li, Na, or K), and the doped α-B sheet is

predicted to be a promising hydrogen storage material [324].

4.4 B2C sheet

B2C sheet is a new B-C phase, proposed by Wu et al. [325], where each C atom

is bonded with four B atoms, forming tetra-coordination C moiety. The B2C sheet is

composed of closely-packed hexagons and rhombi, and its structural design is

essentially the 2D extended network of a predicted ptC molecule CB4 (Fig. 33a). The

B2C sheet has a C2-rotation symmetry and two reflection symmetries with respect to

the vector a and b, respectively. The optimized lattice constants are a = 2.558 Å and b

= 3.453 Å, and the B-C and B-B bond length is 1.557 and 1.685 Å, respectively (Fig.

33b). Note that not forming a completely planar structure, B2C sheet is slightly

corrugated, where B-layer and C-layer are separated by a tiny distance of ~0.085 Å

(Fig. 33c). From the deformation and total electronic density (Fig. 33d and e), the C

Page 85: Graphene-analogous low-dimensional materials

84

atoms form multi-center electron-deficient covalent bonds with four B atoms, and

meantime, the two-center B-B bonds form between two neighboring motifs.

Fig. 33. (a) CB4 molecule (left) and a C2V-CB4 motif (right). (b) Top and (c) side view

of B2C monolayer. (d) Deformation electronic density of B2C. Blue and yellow region

refers to electron rich and deficient region, respectively. (e) Total electronic density

projected on B2C surface. Reproduced from Ref. [325]. Copyright © 2009, American

Chemical Society.

B2C has many appealing properties. For example, the B2C network is predicted

to have excellent mechanical and chemical stability due to strong covalent bonds. The

elastic strength of B2C is found to lie between those of graphene and boron monolayer.

Electronically, B2C sheet and derived nanoribbons are predicted to be uniformly

metallic. Based on first-principles lattice dynamics and electron-phonon coupling

computations, Dai et al. [326] showed that B2C sheet is a 2D phonon-mediated

superconductor with a relatively high transition temperature (19.2 K). Similar to BC3,

B2C sheet can also hold metal atoms strongly, and the Li-doped [327] B2C sheet is a

promising system to storage hydrogen.

Page 86: Graphene-analogous low-dimensional materials

85

5. Non-planar materials

5.1 Metal dichalcogenides

Transition metal dichalcogenides (TMDs), in a general formula MX2 (M = Mo,

W, V, Nb, Ta, Ti, Zr, Hf, and X = S, Se, Te), represent an intriguing family of

materials with prospects for a broad range of unique properties and applications [328].

TMD materials have layered structures consisting of stacks of X-M-X sandwiches

held together by vdW force. Depending on the different compositions, TMDs can

have hexagonal or rhombohedral symmetry, and the central metal atoms can form

octahedral or trigonal prismatic coordination configuration. Due to their weak

interlayer interactions, isolated sheets of TMDs can be cleaved along the layer plane

similar to the case of graphite, and these mono- or few-layer structures are regarded as

non-planar graphene analogues.

Among the large group of layered TMDs, MoS2 and WS2 are the two most

widely studied materials in recent years, and hence in the subsequent section, we

mainly summarize recent advances in experimental fabrications, fundamental

properties, and potential applications of MoS2 and WS2 thin nanosheets/ribbons.

5.1.1 MoS2 and WS2

MoS2 and WS2 exhibit similar structural features. They are both characterized as

hexagonal layered configurations in which the atoms in the layer are bonded with

strong covalent bonding, while the layers are packed together with weak interlayer

forces like graphite and h-BN (Fig. 34). Within MoS2 or WS2 monolayer, each Mo or

Page 87: Graphene-analogous low-dimensional materials

86

W atom is coordinated to six S atoms, and each S atom is coordinated to three Mo or

W atoms, forming a trigonal prismatic coordination configuration.

Fig. 34. Layered structure of MoS2, and its single layer has a thickness of 6.5 Å.

Reproduced from Ref. [328]. Copyright © 2011, Nature Publishing Group.

Experimentally, thin nanosheets of MoS2 and WS2 can be prepared through

many techniques. For example, because of the weakly-bonded layers, single and few

layers of MoS2 and WS2 can be extracted by using top-down techniques such as

micromechanical cleavage, Li-ion intercalation and exfoliation, and liquid-phase

exfoliation (see Sections 2.1 and 2.2). Surface ripples are observed in mechanically

exfoliated single-layer MoS2, like the case of graphene, which reflects similar stability

mechanism of both 2D materials and can explain the degradation of electron mobility

of MoS2 due to exfoliation [329]. Alternatively, using bottom-up technologies, such

as CVD growth on insulating substrates, hydrothermal reaction at high temperature,

and thermal evaporation technique, high-quality MoS2 and WS2 flakes with hundreds

of nanometers to a few microns in size can be generated (see Sections 2.3 and 2.6).

These developed methods also hold great promise for preparing atomically thin films

of other TMDs (such as VS2, TiS2, NbSe2, WSe2, MoSe2, and MoTe2), and other types

Page 88: Graphene-analogous low-dimensional materials

87

of metal chalcogenides (such as GaS, GaSe, Bi2Se3, and Bi2Te3), which has been

discussed in Sections 2.1-2.6.

5.1.1.1 Electronic properties

MoS2 and WS2 sheets. Unlike the semimetallic characteristic of graphene, both

MoS2 [330] and WS2 [331] sheets have intrinsic band gaps, and might serve as an

important complement to graphene in the field of semiconducting applications.

Theoretical studies on the lattice dynamics [332], electronic structures [333], and

dielectric properties [334] of MoS2 have been reported.

The electronic properties of MoS2 are highly associated with their thickness

owing to perpendicular quantum confinement effect (Fig. 35). Bulk MoS2 is an

indirect-band-gap semiconductor (~1.2 eV), and recent first-principles computations

have revealed that intercalation of K atoms into interstitial sites of bulk MoS2 can lead

to a metallic characteristic as a result of electron donation from K 4s orbital to the

conduction band of MoS2 [335]. Bilayer MoS2 is also an indirect semiconductor with

a gap of ~1.6 eV. However, single-layer MoS2 is a direct-band-gap semiconductor

with a larger gap (~1.9 eV) [336,337] and shows a strong photoluminesence absent in

the bulk form [338,339]. The confinement induced indirect-to-direct gap transition in

the single-layer limit also appears in WS2 [331] and other TMDs such as MoSe2,

MoTe2 [340], and WSe2 [341]. The indirect-to-direct gap crossover can also be

realized in few-layered MoSe2 by thermal excitement [342].

Page 89: Graphene-analogous low-dimensional materials

88

Fig. 35. (a) Band structures of bulk and monolayer MoS2. (b) Band structures of bulk

and monolayer WS2. Reproduced from Ref. [337]. Copyright © 2011, American

Physical Society.

The presence of a direct band gap in single-layer MoS2 makes it promising for

integration of MoS2-based microelectronics into photoelectronic devices which allow

for efficient electron-phonon energy conversion. More importantly, the intrinsically

moderate band gap of single-layer MoS2 in the visible frequency range overcomes the

drawback of either zero band gap of graphene or wide band gap of h-BN, implying its

prospective semiconducting applications in complementing graphene or BN as

next-generation nanoelectronics and photoelectronics.

Also, due to its unique electronic structure, MoS2 monolayer has great potential

for valleytronics and spintronics applications. MoS2 monolayer is a two-valley

Page 90: Graphene-analogous low-dimensional materials

89

direct-gap semiconductor where the conduction band and valence band edges are both

located at the K point. Note that the two inequivalent valleys are separated in the

Brillouin zone by a large momentum, the valley index is thus robust against

intervalley scattering, and consequently can be used as a potential information carrier.

Monolayer MoS2 is known to lack inversion symmetry, which can lead to valley Hall

effect and valley-dependent optical selection rules for interband transitions at the

k-points. In addition, compared with the nearly vanishing spin-orbit-coupling in

graphene, MoS2 has a stronger spin-orbit-coupling interaction originated from the d

orbitals of the heavy metal atoms, which results in a strong spin-orbital-induced band

splitting. The inversion symmetry breaking along with the strong spin-orbit

interaction leads to coupled spin and valley physics in MoS2 monolayer [343].

Actually, experimental evidence for the valley polarization of MoS2 monolayer has

been achieved by photoexciting the samples with circularly polarized light, as has

been recently reported by three experimental groups [344-346], and this might lead to

development of potential “valleytronics” devices based on MoS2 monolayer.

Moreover, due to strong in-plane Mo-S covalent bonds, MoS2 nanosheets have

good mechanical strengths. The high Young’s modulus obtained for suspended

ultrathin MoS2 nanosheets with thickness of 5-25 layers (E = 0.33 ± 0.07 TPa) are

comparable to that of graphene oxide [347], and the in-plane stiffness (180 ± 60 Nm-1)

and effective Young’s modulus (270 ± 100 GPa) of single-layer MoS2 are estimated to

be comparable to those of steel [348]. These excellent mechanical properties make

MoS2 nanosheets attractive as reinforcing elements in composites or for fabrication of

Page 91: Graphene-analogous low-dimensional materials

90

flexible electronic devices.

Note that the electronic structures of MoS2 sheets can not only be tuned by

varying the number of layers, but also can be modulated by application of mechanical

strain, electric field, surface adsorption, and defects. For example, Scalise et al. [349]

showed that the band gaps of monolayer and bilayer MoS2 decrease gradually upon

application of bi-axial strain, and finally change to be metallic. Similarly,

semiconductor-to-metal transition also appears for monolayer and bilayer MoSe2,

MoTe2, WS2, and WSe2 under tensile and shear strains [350], or bilayer MoS2, MoSe2,

MoTe2, and WS2 under perpendicular electric field [351]. For the adatom adsorption,

He et al. [352] showed that H, B, C, N, O, and F atoms chemically adsorb on MoS2

monolayer, and these atoms significantly modify the electronic structures into n-type

or p-type semiconductors and induce magnetism in MoS2. The MoS2 sheets can also

obtain local magnetism through adsorption of 3d transition metals (such as Co, Cr, Fe,

Mn, Mo, Sc, Ti, and V), as well as Si and Ge [353].

Atomic defects such as vacancy, dislocation, and doping also serve as significant

modifications for MoS2. Like the defect formation in graphene and BN, atomic

defects in TMDs can be generated by electron irradiation. Specifically, Liu et al. [354]

observed the W and S monovacancies in monolayer WS2 nanoribbons encapsulated

inside carbon nanotubes at an electron acceleration voltage of 60 kV. The vacancies

were created mainly at the ribbon edges, and the W vacancy was found to induce a

drastic structural deformation. Moreover, Hansen et al. [355] observed the atomic Mo

(S) defects at the edges of MoS2 clusters under an 80 kV electron irradiation. Komsa

Page 92: Graphene-analogous low-dimensional materials

91

et al. [356] have recently reported the vacancy formation in MoS2 sheets under an 80

keV electron beam. The HRTEM image in Fig. 36a clearly indicates formation of a

number of S vacancies accompanied by the crack formation and lateral shrinkage of

the sheet. The simulated TEM images of single and double S vacancies (Fig. 36 d and

e) are in consistency with the observed vacancy structures in measured TEM images

(Fig. 36f). Atomic vacancies can be further filled by other impurity species, forming

electron-beam-mediated substitutional doping defects.

Fig. 36. (a) HRTEM image of single-layer MoS2. Models and simulated HRTEM

images for (b, d) single and (c, e) double vacancy, respectively. Experimental TEMs

are shown in (f). Reproduced from Ref. [356]. Copyright © 2012, American Physical

Society.

Theoretical studies indicated that creation of monovacancy (Mo or S) or

divacancy (2S or MoS) defect in MoS2 monolayer does not induce any magnetic

moment, while the construction of MoS2 triple vacancies gives rise to significant

magnetic moments [353]. The MoS2 triple vacancy in MoS2 is predicted to be able to

attract H2O molecule and catalyze the dissociation of H2O into its constituent H and O

atoms [357].

Page 93: Graphene-analogous low-dimensional materials

92

In a very recent study, Zou et al. [358] have investigated the dislocations and

grain boundaries in single layers MoS2 and WS2. Their simulations indicated that the

dislocation cores in MoS2 and WS2 form concave polyhedras (5|7, 6|8, and 4|6

polygons) with dreidel-like shapes. The linear arrays of these dislocations then form

grain boundaries between the mutually tilted crystalline domains. The presence of

grain boundaries introduces midgap electronic states within the band gap of MoS2 or

WS2 and acts as sinks for the carriers.

As for doping defects, most theoretical investigations focused on mixed TMD

phases with either metal or chalcogen mixing. Ivanovskaya et al. [359] studied the

bilateral doping within the mixed MoS2-NbS2 system, and found that substitutional

Nb doping in MoS2 leads to a semiconductor-metal transition, while Mo doping in

NbS2 does not alter the metallic character of NbS2 system. On the other hand, Komsa

et al. [360] studied the single layers of MoS2xSe2(1−x) alloy. The lowest-energy

configurations of MoS2xSe2(1−x) favor having dissimilar atoms in the nearest neighbor

sites in the chalcogen sublattice. The band structures of the alloy systems resemble

their binary MoS2 or MoSe2 constitutes, and their direct gaps (1.65-2.0 eV) can be

tuned between the constituent limits. The mixed TMD materials can be prepared by

CVD growth or be exfoliated from bulk mixed phases, and their tunable properties

should be beneficial for optoelectronic applications.

MoS2 nanoribbons. Recently, MoS2 derived nanoribbons have also fascinated the

research community. Li et al. [28] computationally studied MoS2 nanoribbons with

widths up to ~6 nm (Fig. 37). Armchair MoS2 nanoribbons are non-magnetic and

Page 94: Graphene-analogous low-dimensional materials

93

semiconducting, and their direct band gaps converge to nearly ~0.56 eV as the width

increases (Fig. 37d). Due to edge effect, the band gaps of armchair MoS2 nanoribbons

are smaller than that of the 2D single-layer MoS2. In comparison, zigzag MoS2

nanoribbons are ferromagnetic metals. The induced magnetism mainly concentrates

on the edge Mo and S atoms (Fig. 37c), and the total magnetic moments increase with

increasing ribbon width or thickness. Note that ferromagnetism has been detected in

flat MoS2 clusters and MoS2 sheets, which might be partly due to the presence of

zigzag edges at the grain boundaries of MoS2 crystals or the presence of lattice defects

[361 ,362]. Zigzag MoS2 nanoribbons are energetically more feasible than the

armchair counterparts with comparable widths, while the case in graphene

nanoribbons is opposite. Many other theoretical investigations have also deepened the

understanding of MoS2 nanoribbons, and similar to MoS2 sheets, the properties of

MoS2 nanoribbons can be modulated by edge functionalization (such as H, and S),

adatom adsorption, and vacancy defects, or by applying strains or electric fields

[363-365].

Fig. 37. Geometries of 8-zigzag (a) and 15-armchair (b) MoS2 nanoribbon. The spin

Page 95: Graphene-analogous low-dimensional materials

94

density distribution of 8-zigzag MoS2 nanoribbon (c), and the energy gap variation of

armchair nanoribbons as a function of ribbon width (d) (8 ≤ Na ≤ 20, inset).

Reproduced from Ref. [28]. Copyright © 2008, American Chemical Society.

Ultra-narrow MoS2 and WS2 nanoribbons with zigzag-shaped edges have been

synthesized by Wang et al. via the bottom-up chemical growths inside carbon

nanotubes (see Section 2.5.2) [21,22]. For these zigzag-edged MoS2 nanoribbons, the

Mo edge tends to be half-saturated by S atoms, while the S edge prefers to be bare.

5.1.1.2 Potential applications

Transistors. The semiconducting ultra-thin MoS2 layers and other TMDs,

because of their intrinsic advantages such as high structural stability, tunable band

gaps between 1-2 eV which can lead to high on/off ratios, subnanometer thickness,

and high carrier mobility, are attractive for use as channel materiasl in low-power

FETs. The TMD-based FETs generally possess a high on/off ratio of larger than 106,

and the carrier mobility of around 200-500 cm2/(Vs) is comparable to the

single-crystal Si [366]. As the first attempt of integrating the mechanically exfoliated

single-layer MoS2 into FET electronics, Radisavljevic et al. [328] fabricated a

MoS2-based n-type transistors using HfO2 as the gate dielectric (Fig. 38). This device

demonstrates a high room-temperature channel mobility of more than 200 cm2/(Vs)

and current on/off ratios of up to 1×108, comparable to the mobility achieved in thin

Si films or GNRs. A high performance p-type FET based on single layered WSe2 with

Page 96: Graphene-analogous low-dimensional materials

95

hole-doped source/drain contacts and high-κ top-gate dielectric (ZrO2) has also been

fabricated, which exhibited a hole mobility of ~250 cm2/ (V s), subthreshold swing of

~60 mV/dec, and on/off ratio of >106 [367]. Note that the coating of high-κ dielectric

plays a significant role in improving the carrier mobility of semiconducting channel

materials due to the dielectric engineering.

Fig. 38. Schematic of HfO2-top-gated monolayer MoS2 FET device. Reproduced from

Ref. [328]. Copyright © 2011, Nature Publishing Group.

Many other groups also reported the high performance of FETs made of MoS2

films based on the mechanical exfoliation technique [368-370]. For example, Zhang

et al. [371] fabricated an electric double layer transistor (EDLT, a FET gated by ionic

liquids) using thin MoS2 flakes as channel materials and demonstrated that this device

displays ambipolar transistor operation. Braga et al. [372] realized an ambipolar FET

by coupling exfoliated thin flakes of WS2 with an ionic liquid dielectric. Kim et al.

[373] showed high-mobility and low-power thin-film FETs based on the multilayer

MoS2 crystals. MoS2 FET devices are sensitive to environmental moisture, and

exposure to ambient oxygen or water can drastically reduce the device performances

[374].

Page 97: Graphene-analogous low-dimensional materials

96

The MoS2/WS2 layers prepared from liquid exfoliation or CVD growths have

also been implemented into FET fabrications, and they exhibit similar electrical

characteristics [52]. For example, Hwang et al. [375] reported the realization of

back-gated FETs made of the chemically exfoliated multilayer WS2, which

demonstrate a high (~105×) on/off current ratio, and show clear photo response to

visible light. Pu et al. [376] fabricated a CVD-grown MoS2 thin-film transistor using

ion gel (gelation of an ionic liquid) as gate dielectrics, and this device is characterized

by a high on/off ratio of 105 and a mobility of 12.5 cm2/(V·s) at a low operating

voltage of 0.68 V.

Moreover, the large in-plane carrier mobility of MoS2 also makes it attractive for

fabrication of other digital devices such as phototransistor, integrated circuits, and

signal amplifier. For instance, Yin et al. [377] fabricated a phototransistor based on

MoS2 single layer, which exhibits better phtoresponsivity (7.5 mA/W) than the

graphene-based device (1 mA/W). Lee et al. [378] fabricated top-gate phtotransistors

based on single-, double-, and triple-layer MoS2 sheets, and their devices with triple

MoS2 layers exhibited excellent photodetection capabilities for red light, while those

with single- or double-layers were useful for green light detection. The first

implementation of single-layer MoS2 into the integrated circuits was reported by

Radisavljevic et al. [379], and this device can be used to operate as inverters. Wang et

al. [380] also fabricated the fully integrated multistage circuits based on bilayer MoS2

transistors, such as a logic inverter, a static random access memory, and a five-stage

ring oscillator.

Page 98: Graphene-analogous low-dimensional materials

97

Sensors. MoS2 nanosheets are also favored in electronic sensing applications.

The EFTs made from MoS2 sheets have been used to detect adsorption of NO gases

[370]. The sensing mechanism is ascribed to the p-type doping effect of the adsorbed

NO molecules, which changes the electrical resistance of the originally n-type doped

MoS2. Likewise, a flexible thin-film transistor using MoS2 as the active channel and

reduced graphene oxide (rGO) film as the drain and source electrodes is used as a gas

sensor for NO2 detection [ 381 ]. Furthermore, the electrochemically reduced

single-layer MoS2 nanosheets have been demonstrated as sensitive glucose and

dopamine detectors [382].

Energy harversting. MoS2 and WS2 nanosheets can be additionally applied to

energy haversting field. One hand, the laterally confined layered structures of MoS2

and WS2 nanosheets favor their use as active electrode materials in Li-ion batteries.

For example, Seo et al. [115] prepared WS2 nanosheets from W18O49 precursor, and

showed that the electrochemical lithiation capacity of WS2 nanosheets was greatly

improved as compared with the bulk WS2, owing to their enhanced surface area and

easier diffusion accessibility of Li ions. Xiao et al. [383] prepared exfoliated

MoS2/polyethylene oxide (MoS2/PEO) nanocomposites and demonstrated high

capacity (1000 mAh g-1) and good cycle stability for the composite. Moreover, by

combining semiconducting MoS2 with conductive carbon materials (graphene, carbon

nanotubes, or amorphous carbon), better electrochemical performances are expected

[384,385]. Zigzag MoS2 nanoribbons were computationally predicted to effectively

bind the Li ions close to the edges with high Li mobility, and can be alternatively

Page 99: Graphene-analogous low-dimensional materials

98

developed as cathode materials for Li-ion batteries [386]. Besides, MoS2 thin layers

are also promising cathode materials for rechargeable Mg batteries [114].

On the other hand, the tunable band gaps of MoS2 and WS2 in the visible regions

make them potential candidates for photovoltaic solar cell applications. A hybrid bulk

heterojunction (BHJ) solar photovoltaic cell employing MoS2/TiO2 nanocomposites

and poly 3-hexylthiophene (P3HT) as the active layers is demonstrated to have a

photoconversion efficiency of 1.3% [387]. A Schottky-barrier solar cell made from

MoS2 nanomembranes, indium tin oxide (ITO), and Au with a stacked structure of

ITO-MoS2-Au has been recently demonstrated to possess a photoconversion

efficiency of 1~2% [388].

5.1.2 Other layered metal dichalcogenides

MoS2 and WS2 only represent two examples of the broad family of layered

TMDs. Aside from them, many other metal chalcogenides have also earned

tremendous attention, such as MoSe2 [389], WSe2 [390], VS2, VSe2 [391], NbS2,

NbSe2 [392], SnS2 [393], Bi2Se3 [394,395], and Bi2Te3 [396]. For example, WSe2

[390] would undergo an isostructural semiconductor-semimetal phase transition under

high pressure. By applying a tensile strain, a ferromagnetic behavior can be

theoretically induced in VS(Se)2 [391] and NbS(Se)2 [392] monolayers. Sun et al.

[393] synthesized SnS2 single layers through liquid exfoliation, and these single layers

have superior photocatalytic performances to the bulk phase in achieving effective

visible-light water splitting. Particularly, Bi2Se3 and its related materials such as

Page 100: Graphene-analogous low-dimensional materials

99

Bi2Te3, and Sb2Te3 were identified as a new class of topological insulators [394],

which were proposed for the realization of dissipation-less interconnects, low-power

electronics and quantum computing devices, as well as for thermoelectric

applications.

5.2 Layered oxide and hydroxide nanosheets

Layered oxides and hydroxides typically have strong interlayer interactions

(mostly electrostatic interactions), and their nanosheets are usually synthesized by

chemical exfoliation of the parent layered crystals in exfoliating solutions. So far, a

wide variety of oxides (such as Cs0.7Ti1.825O4, Ca2Nb3O10, K0.45MnO2, K4Nb6O17,

RbTaO3, KTiNbO5, and Cs6+xW11O36) and hydroxides (such as

M2+1−xM3+

x(OH)2·An−x/n mH2O, where M2+ = Mg2+, Fe2+, Co2+, Ni2+, Zn2+, etc., and

M3+ = Al3+, Fe3+, Co3+; RE(OH)2.5 xH2O·An−0.5/n, where RE = Nd, Sm, Eu, Gd, Tb, Dy,

Ho, Er, Tm, Yb, Lu, Y, etc.) have been delaminated into their charge-bearing 2D

nanosheets (see the recent review [397]). The electrostatic repulsion between the

negatively charged oxide nanosheets or the positively charged hydroxide nanosheets

leads to the formation of highly stable aqueous colloidal solutions.

Generally, the exfoliation of most layered metal oxides is achieved firstly by the

protonation reaction of pristine oxides through acid treatment (such as HCl), and

followed by the intercalation of organic ions into the protonated oxides via the

addition of chemical intercalators. The most commonly used intercalator is

tetrabutylammonium ion (TBA+), although tetramethylammonium ion (TMA+) and

Page 101: Graphene-analogous low-dimensional materials

100

ethylammonium ion have also been successfully used. The intercalation process of

organic ions (such as TBA+) was mediated via ion-exchange reactions (such as

H+-TBA+), which expands the host structures and decreases the electrostatic

interactions between host layers, assisting the exfoliation of layered materials. In a

similar way, layered metal hydroxides can be exfoliated through insertion of bulky

intercalated agents (such as NO3−, ClO4

−, and dodecyl sulfate (DS) ions) and

subsequent treatment with organic solvents (such as butanol, and formamide) under

stirring, heating, or sonicating conditions. As an example, Fig. 39 presents the

schematic illustration for the exfoliation of DS ion intercalated nickel hydroxide [398].

In this process, exfoliation was induced by intercalation of DS ions into Ni(OH)2

under the reflux of formamide solution. AFM images show homogeneous thickness of

1.1 nm for the exfoliated hexagonal layers, larger than that estimated from

crystallographic thickness, which is likely due to surface absorption of water and DS

ions.

Fig. 39. Schematic for exfoliation of DS ion intercalated nickel hydroxide and AFM

image of the isolated hexagonal layers. Reproduced from Ref. [398]. Copyright ©

2008, American Chemical Society.

Page 102: Graphene-analogous low-dimensional materials

101

2D oxide and hydroxide nanosheets exhibit diverse properties depending on their

compositions, and demonstrate many promising applications. Most oxide nanosheets

are wide band-gap semiconductors, thereby facilitating their potential use as

semiconducting hosts, photocatalysts, dielectric materials, etc. For example, titanium

oxide nanosheets present excellent photochemical properties [399], and the Co

substituted titanium oxide nanosheets can achieve room-temperature ferromagnetism

[400]. Nanosheets of MnO20.45- [401] and RuO2.1

0.2- [402] undergo electrochemical

redox reactions and are useful for electrochemical energy storage. The Cs4W11O362-

nanosheets exhibited photochromic properties, showing reversible color change upon

UV irradiation [403]. In the case of double hydroxide nanosheets, due to their unique

anion-exchangeability and biocompatibility, they are useful candidates as anion

exchangers to remove toxic anions or as drug and gene delivery nanocarriers for

pharmaceutical and biological applications [404]. Moreover, hydroxide nanosheets

can be used as catalysts or catalyst supporters. For example, the tungstate-exchanged

layered double hydroxide is a promising biomimetic catalyst for oxidative

bromination [405]. In addition, as in the case of oxide nanosheets, hydroxide

nanosheets also exhibit versatile electronic, magnetic, and optical properties, and most

of them have ferromagnetic and half-metallic characteristics. For example, Co2+-Al3+

hydroxide nanosheets act as nanoscale ferromagnetic layers and exhibit an interesting

magneto-optical response in the UV-visible region [406]. Furthermore, nanosheets of

europium hydroxide exhibit characteristic red emission based on the Eu3+

photoactivator, which suggests potential applications of ultra-thin films of this

Page 103: Graphene-analogous low-dimensional materials

102

material as optical devices [407].

In addition to the layered oxides and hydroxides with strong electrostatic

attractions between the host layers and the cationic (for oxides) or anionic (for

hydroxides) interlayer species (as discussed above), we will consider another

exceptional layerd oxide with weakly bonded layers, V2O5, in the subsequent part.

Fig. 40. Structure model of bulk V2O5.

Bulk V2O5 has an orthorhombic symmetry, and is a typical oxide

semiconductor with a layered structure stacking along the (001) direction,

characterized as strong in-plane V-O covalent bonding and weak vdW coupling

between adjacent layers (Fig. 40) [408]. The individual V2O5(001) monolayer is

strongly corrugated and built up of distorted VO5 square pyramids, which are

alternatively pointed upward and downward. Three inequivalent O atoms exist on

the monolayer: the outermost single-coordinated O atoms (O1) are doubly bonded to

the V atoms, while the other two types are two- and three-coordinated bridging O

atoms (O2 and O3). Bulk V2O5 has a visible band gap of ~2.4 eV, and the (001)

monolayer has similar physical properties and stability as its bulk crystal and can be

in principle produced by cleaving the weak interlayer interaction. As a matter of fact,

the thin nanosheets of V2O5 have been recently fabricated via liquid exfoliation

Page 104: Graphene-analogous low-dimensional materials

103

technique in polar solvents [48].

Fig. 41. (a) Schematic of exfoliation of layered bulk V2O5 into {001}-oriented

few-layer V2O5 nanosheets. (b-d) Cathode performance of {001}-oriented few-layer

V2O5 nanosheets and bulk V2O5 in Li-ion batteries: (b) initial galvanostatic

charge–discharge voltage profiles at a current density of 59 mA g-1 (0.2 C). Here, 1 C

= 294 mA g-1; (c) cycling performance at 0.2 C; (d) rate capability at various charge

and discharge rates. (e) The Ragone plot of V2O5 nanosheets for power and energy

performance, in comparison with some advanced energy storage and conversion

devices. Reproduced from Ref. [48]. Copyright © 2013, Royal Society of Chemistry.

Page 105: Graphene-analogous low-dimensional materials

104

The high surface-bulk ratios and the absence of interlayer interactions offer 2D

V2O5 nanosheets unique electronic properties and important functional applications.

According to DFT computations, the single-layer V2O5 sheets and nanoribbons

possess tunable electronic and magnetic properties that are susceptible to surface

hydrogenation [409]. Du et al. [410] pointed out that V2O5 monolayer could

theoretically promote the dehydriding kinetics in magnesium hydride. Recently, Rui

et al. [48] have successfully prepared few-layer V2O5 nanosheets with thickness of

3.1-3.8 nm through direct exfoliation of bulk V2O5 in formamide (Fig. 41). Because

of the short Li diffusion paths provided by the ultrathin thickness, the obtained V2O5

nanosheets demonstrated excellent performances as cathode materials in Li-ion

batteries. As shown in Fig. 41b-d, the ultrathin V2O5 nanosheets display larger

reversible Li capacity, higher Coulombic efficiency, and better rate capability than the

bulk V2O5 electrode. Moreover, the V2O5 nanosheet electrode achieves a higher

energy density (158 W h/kg at a power rate of 20 kW/kg) than the pseudocapacitors

(1–70 W h/kg at a power rate of 1-20 kW/kg) (Fig. 41e), indicating the possible

applications of V2O5 nanosheets as superior electrochemical energy-storage devices

with both high-power and high-energy densities.

5.3 Graphitic-like phase of ZnO

Most of the group II-VI and III-V binary compounds such as ZnO, AlN, and

GaN are naturally crystallized in a stable wurtzite structure with a diamond-like

configuration (Fig. 42 a). However, first-principles computations [411,412] predicted

Page 106: Graphene-analogous low-dimensional materials

105

that the polar (0001)-oriented ultrathin films of wurtzite materials (such as AlN, BeO,

GaN, ZnO, and ZnS) with only a few atomic layers favor a new form of stable

graphitic structures, where each cation and anion adopt a sp2 hybridized state and are

arranged in planar three-fold coordination instead of the bulk-like tetrahedral

configuration (Fig. 42b,c). This interesting structural transformation is driven by

surface charge transfer mechanism, and the surface polarity is completely removed

after forming the graphitic-like structure. Unlike the metallic characteristic of polar

(0001) wurtzite films, the depolarized graphitic structures are semiconducting instead.

Fig. 42. (a) Wurtzite structure of bulk ZnO. Zn atoms are depicted in light blue, O in

red. Reproduced from Ref. [411]. Copyright © 2005, Royal Society of Chemistry. (b)

Polar wurtzite structure of ZnO(0001) film. (c) Graphitic structure. Reproduced from

Ref. [412]. Copyright © 2006, American Physical Society.

Shortly after the appealing theoretical prediction, the presence of graphitic phase

was confirmed by an epitaxial-growth experiment of ultrathin ZnO films. Tusche et al.

[413] observed two-monolayer-thick non-polar ZnO(0001) films on Ag substrates,

which provided a direct evidence for the presence of planar ZnO sheets (Fig. 43). The

observed depolarization of two-monolayer ZnO film is accompanied by a lateral 1.6%

Page 107: Graphene-analogous low-dimensional materials

106

expansion of the lattice parameter and a 3% reduced Zn-O bond length within the

sheets. The experimentally observed graphitic ZnO films on Ag substrate can only

persist up to 2-3 layers thickness, and then a transition to the bulk wurtzite takes place,

which is much lower than the theoretical predicted 16 layers. This inconsistency can

be explained by the possible effects such as structure defects, film roughening, and

compressive strain induced by lattice mismatch. Noteworthy, except for graphene-like

ZnO observed in experiments, other types of graphene-like II-VI (ZnS, BeO, or MgO)

and Ш-V (AlN, or GaN) phases still remain hypothetical.

Fig. 43. Models of wurtzite-like ZnO(0001) bilayer structure (a) and the

experimentally determined graphitic-like structure (b). High resolution STM image of

graphite-like ZnO film about 2.2 monolayers in thickness. Reproduced from Ref.

[413]. Copyright © 2007, American Physical Society.

The electronic and magnetic properties of ZnO [414], GaN [415,416], AlN [417],

BeO [418] and MgO [419] in their graphitic hexagonal phases have been intensively

examined, and we mainly discuss the case of ZnO since similar properties are

expected in other types of graphitic phases. The ZnO monolayer is a direct-gap

semiconductor (2.62 eV at the DFT/GGA level, and 3.57 eV under the GW correction

Page 108: Graphene-analogous low-dimensional materials

107

[420]), and presents ferromagnetic coupling after Co-doping [421]. The creation of

Zn vacancy brings on local magnetic moments, while other types of defects such as O

vacancy, Zn+O double vacancy, and Zn-O antisite defects cause no magnetism for the

system [414]. Doping the O sublattice of ZnO monolayer with C or B atoms results in

a “NM→anti-FM→FM” transition, accompanied by a semiconductor→half metal

transition, and doping by N atoms leads to a p-type semiconductor with the anti-FM

ground state [422]. Moreover, the electronic and magnetic properties of graphene-like

ZnO sheets can be modulated by adatom adsorption of H and F atoms [423,424]. For

example, DFT computations on hydrogenation of graphitic ZnO nanosheets revealed

that the electronic and magnetic properties depend on the adsorption sites of H sites

and the sheet thickness.

For the cutted 1D ribbons of ZnO, the single-layered armchair ZnO nanoribbons

are non-magnetic semiconductors, and their band gaps tend to reduce monotonically

with increasing transverse field strength [425]. The single-layered zigzag ZnO

nanoribbons are metallic and exhibit ferromagnetic responses, which emerge from the

unpaired electrons localized on the oxygen-terminated edges. The edge magnetism of

zigzag ZnO nanoribbons can be modulated by applying an external field [426] (Fig.

44) or edge functionalization with sulfur or thiol groups [427]. In addition, half

metallicity was found in edge-passivated wide ZnO nanoribbons when only Zn edge

is saturated by hydrogen, CH3- or NH2- groups [428]. By applying a proper

axis-strain, one can induce a metal→semiconductor transition in zigzag ZnO

nanoribbons [429]. Interestingly, Botello-Méndez et al. [430] found that in multilayer

Page 109: Graphene-analogous low-dimensional materials

108

zigzag ZnO nanoribbons, edge atoms on top layer form covalent bonds with edge

atoms on bottom layer. As a result, the magnetism and metallicity disappear in the

even-numbered layer. The metallicity still retains in odd-numbered layer but the

magnetic moment quickly vanishes as the layer number increases to five layers.

Fig. 44. (left) Schematic of a zigzag ZnO nanoribbon under a transverse electric field.

(right) Magnetic moments per supercell of different zigzag ZnO nanoribbons as a

function of applied electric fields. Reproduced from Ref. [426]. Copyright © 2010,

American Chemical Society.

5.4 MXenes

Another family of layered materials, the layered ternary metal carbides, nitrides

or carbonitrides, also known as “MAX” phase, has fascinated the researchers’

interests quiet recently. The MAX phases have a general formula of Mn+1AXn (n=1, 2,

3), where M represents early transition metals (M=Ti, Sr, V, Cr, Ta, Nb, Zr, Mo, or Hf),

A represents main-group sp elements (mostly III A or IV A), and X represents either C

or N as well as both. They correspond to a big family with more than 60 members and

constitute a new class of layered materials combining the properties of both metals

Page 110: Graphene-analogous low-dimensional materials

109

and ceramics [431].

Structurally, MAX phase can be described as the inter-growth structures with

alternative stacking of hexagonal MX layers and close-packed planar A atomic layers.

As an example, Fig. 45 presents a typical structure of M2AX systems, where the

chemical bonds between A-M elements are weaker than the covalent M-X-M bonds,

which makes it possible to extract A layer from the layered solid. However, different

from graphite with weak vdW interlayer interaction, the MX layers in MAX phase are

held together by A-containing layers via partially ionic bonding, and the consequential

bond strengths are quite strong, making separation of MX layers from MAX phases

cannot be easily realized with mechanical cleavage or direct dispersion and

ultrasonication.

Fig. 45. Typical bulk structure of layered M2AX, where M, A, and X represent

transition metal, A-element, and carbon/nitrogen, respectively.

To isolate elementary MX layers from bulk MAX, one alternative strategy is to

use an effective reactant to selectively etch the interleaved A-containing layers to

decrease the interlayer interactions, but without destroying the layered morphology of

Page 111: Graphene-analogous low-dimensional materials

110

MX layers. A pioneering work in this line by Gogotsi group is the successful

exfoliation of Ti3C2 nanosheets from Ti3AlC2 by firstly immersing Ti3AlC2 powers

into 50% HF (Ti3AlC2 + 3HF = AlF3 + 3/2H2 + Ti3C2) to extract the Al layer, and

followed by ultra-sonication treatment in methanol (Fig. 46) [432]. TEM images show

that the exfoliated Ti3C2 nanosheets exhibit hexagonal symmetry, with thickness

ranging from single layer and double layers, to multilayers. Such as-designed

procedure by Gogotsi group has been successfully applied to isolate other MX layers,

Ti2C, Ta4C3, TiNbC, (V0.5Cr0.5)3C2, and Ti3CNx (x < 1), starting from their

Al-containing MAX phases [433]. Under the aqueous environment of HF solutions,

the outer surfaces of the exfoliated MX layers are usually chemically-terminated with

F or/and OH functional groups. With the ease of the as-developed method, one can

assume that more and more MX nanosheets could be exfoliated from their bulk MAX

phases.

Fig. 46. (a) Exfoliation process for Ti3AlC2. (b) TEM images of exfoliated MXene

nanosheets. Inset shows selected area electron diffraction (SAED) pattern confirming

Page 112: Graphene-analogous low-dimensional materials

111

hexagonal symmetry of the planes. (c) TEM images of single- and double-layer

MXene sheets. Reproduced from Ref. [432]. Copyright © 2011, Wiley-VCH.

These isolated 2D carbide and carbonitride nanosheets (Mn+1Xn stoichiometry)

are alternatively termed as “MXene”, as suggested by Gogotsi group, since they have

hexagonal structure and electronic properties comparable to those of graphene.

MXene materials display diversified electronic properties depending on the

appropriate surface treatment. For example, bare Ti3C2 monolayer acts as a magnetic

metal, while its derived Ti3C2F2 and Ti3C2(OH)2 are semiconductors with small band

gaps [434]. Khazaei et al. have recently studied the electronic properties of various

MXene systems, M2C (M = Sc, Ti, V, Cr, Zr, Nb, or Ta) and M2N (M = Ti, Cr, or Zr)

with surface chemically functionalized with F, OH, or O groups [435]. All the bare

MXene monolayers are electrically metallic. After surface functionalization, the

Sc2CF2, Sc2C(OH)2, Sc2CO2, Ti2CO2, Zr2CO2, and Hf2CO2 become semiconductors

with band gaps around 0.25-2.0 eV, and the functionalized Cr2CF2, Cr2C(OH)2,

Cr2NF2, Cr2N(OH)2, and Cr2NO2 become ground-state ferromagnetic.

The metallic or narrow-band-gap semiconducting characteristics endow the

MXene-based materials with intrinsic advantage of good electrical conductivity, and

hence favor their potential application as electrode materials for Li-ion batteries. A

theoretical study on Ti3C2 monolayer shows that Li adsorption forms strong

interaction with Ti3C2 host, and the bare Ti3C2 monolayer exhibits a low barrier (0.07

eV) for Li diffusion and high Li storage capacity (up to Ti3C2Li2 stoichiometry) [434].

Page 113: Graphene-analogous low-dimensional materials

112

This gives insightful prospect for the application of Ti3C2 nanosheets as an anode

material for Li ion batteries. Surface functionalization by F or OH groups, however,

degrades the Li diffusion and decreases the Li storage capacity, and thus should be

avoided in synthetic experiments. The performance of MXene as an anode material

has also been tested in the experiments. Naguib et al. [436,437] conducted detailed

electrochemical tests to evaluate Ti2C’s performance as an anode material in Li-ion

batteries designed for rapid charging and discharging. The key finding in their study

was that the material retained its charge capacity during 1,000 rapid

charging/discharge cycles.

6. Two-dimensional coordination and covalent organic polymers

Linking molecular building blocks into extended networks by either coordination

bonding or strong covalent bonding is now commonly used in the synthesis of

coordination polymers or covalent organic polymers. These materials, especially their

2D forms, give a novel class of atomically regular and uniform 2D crystals beyond

graphene, and exert great scientific attention for both their synthesis and applications.

6.1 Coordination polymers

Coordination polymers (CPs) are a family of nanoporous materials constructed

by organic linkers and metal centers. Depending on the selective building blocks

(metals and ligands), they exhibit a rich variety of architectures and diverse

physicochemical properties. There are numerous examples of coordination polymers

obtained in crystal phase as lamellar materials. However, 2D coordination networks

Page 114: Graphene-analogous low-dimensional materials

113

created as ordered monolayers are rarely reported.

Encouragingly, impressive progress has been achieved in isolating

one-atom-thick CP flakes. Amo-Ochoa et al. [15] presented successful exfoliation of

single layer CP extracted from layered crystals of [Cu2Br(IN)2]n (IN = isonicotinato)

using liquid-phase sonication. AFM topography image (Fig. 47b) shows a very dense

and homogenous distribution of [Cu2Br(IN)2]n flakes on highly oriented pyrolitic

graphite (HOPG) substrate. The typical heights (Fig. 47c) of 5 ± 0.15 Å are in good

agreement with the thickness (5.5 Å) expected for single atomic layers. Theoretical

investigations [438] verified that monolayer [Cu2Br(IN)2]n has a strong tendency to

adsorb NO and NO2, which suggests the technological possibility of using

[Cu2Br(IN)2]n CP as molecular sensors.

Fig. 47. (a) Schematics of [Cu2]3+ coordination environment, single layer

[Cu2Br(IN)2]n, and the bulk crystal of [Cu2Br(IN)2]n. (b) AFM topography image of

sonicated [Cu2Br(IN)2]n on HOPG. (c) Height of profile across the green line in (b).

Reproduced from Ref. [15]. Copyright © 2010, Royal Society of Chemistry.

In addition to exfoliating CP nanosheets from layered crystals, a direct

self-assembly is a more straightforward procedure to create 2D coordination polymers,

Page 115: Graphene-analogous low-dimensional materials

114

and co-deposition of metals and organic ligands on crystalline solid surface offers a

promising approach. Li et al. [ 439 ] reported self-assembly of

5,10,15,20-tetra(4-pyridyl)porphyrin (2HTPyP) with Cu on Ag substrate (Fig. 48a). In

this process, Cu reacts in two ways: for one way, neutral Cu(0) atoms are coordinated

by peripheral pyridyl (py) substituent to form py-Cu-py bridge coordination; for

another way, Cu(0) is coordinated by inner nitrogen atoms of the porphyrin

macrocycle and, at elevated temperatures (450 K), oxidized to Cu(II). This

coordination network represents a mixed-valence polymer consisting of an ordered

array of Cu(II) and Cu(0) centers (Fig. 48b).

Fig. 48. (a) Self-assembly of 2HTPyP and Cu to form a mixed-valence Cu(II) and

Cu(0) network. (b) High-resolution image of 2HTPyP with Cu deposited on Au(111)

surface after 450 K annealing: yellow refers to Cu(II) and green to Cu(0). Reproduced

from Ref. [439]. Copyright © 2012, American Chemical Society.

Moreover, Abel et al. [16] reported formation of polymeric Fe-phthalocyanine

(poly-FePc) single sheet, which was obtained by co-evaporation of Fe and

1,2,4,5-tetracyanobenzene (TCNB) in 1:2 stoichiometry on surfaces such as Ag(111),

Au(111), as well as the insulating NaCl/Ag(100) surfaces. The poly-FePc film has a

Page 116: Graphene-analogous low-dimensional materials

115

square structure with a measured periodicity of 1.15 ± 0.1 nm (Fig. 49). The

poly-FePc nanosheet is theoretically predicted to be an anti-FM semiconductor,

whereas a FM half-metal can be achieved when the central Fe atoms are replaced by

Mn [440].

Fig. 49. (a) Reaction scheme of poly-FePc. (b,c) STM images of poly-FePc formed on

Ag(111) surface. (d) STM image of poly-FePc formed on insulating NaCl island

deposited on Ag(100). Reproduced from Ref. [16]. Copyright © 2010, American

Chemical Society.

Besides the metal-porphyrin complexes, other types of organometallic sheets

have been recently reported. For example, Bauer et al. [ 441 ] synthesized a

monolayered organometallic sheet based on the reversible complexation of Fe2+ and

hexafunctional terpyridine monomers. Shi et al. [442] reported that on Au(111),

molecular ligands of 1,3,5-tris(pyridyl)benzene (TPyB) can form

metallo-supramolecular self-assembly with either Cu or Fe. Beyond the experimental

reports, Kan et al. [ 443 ] theoretically designed organometallic porous sheets

assembled by transition metals (Mn and V) and benzene molecules, which are ideal

magnetic materials with room-temperature ferromagnetism.

Page 117: Graphene-analogous low-dimensional materials

116

6.2 Covalent organic polymers

Covalent organic polymers or frameworks (COFs), made entirely from light

elements (H, C, N, B, and O), are another intriguing class of porous materials that

allow for the integration of organic molecular units with strong covalent bonds into

rigid 2D or 3D architectures. Compared with coordination polymers, COFs provide a

great technological benefit concerning the aspect of their higher structural stability,

since the intermolecular covalent bonds are more stable than the coordination bonds.

Fabrications of COFs typically involve solvothermal reactions of

self-condensation of boronic acids or co-condensation with polyols [444], imine bond

formation [445], trimerization of nitriles [446], borosilicate bond (B-O-Si) [447] and

hydrazone bond formation [ 448 ]. In addition, surface-confined polymerization

reactions of 2D COFs have gained extensive interests, such as radical addition [449],

surface condensation reaction [450], and STM tip- and electron beam-induced surface

polymerization [451].

Most of the produced COF crystals, the boronate-ester derived COFs in

particular, are layered structures composed of π-π stacked covalent sheets that exhibit

staggered graphite-like or elipsed BN-like arrangements, and the single layer sheet

represents a promising model for 2D COFs. Recently, the oriented COF multilayer

films (hexagonal COF-5) grown on graphene substrate have been demonstrated by

solvothermal condensation of 1,4-phenylenebis(boronic acid) (PBBA) and

2,3,6,7,10,11-hexahydroxytriphenylene (HHTP), which showed improved crystallity

than the COF powders (Fig. 50) [452]. The COF-5 monolayer is predicted to be a

Page 118: Graphene-analogous low-dimensional materials

117

semiconductor with a band gap of 4.0 eV (HSE functional), and its electron and hole

carriers are separated into the HHTP and PBBA units, respectively, suggesting the

potential applications of COF-5 in organic electronic and photovoltaic devices [453].

Fig. 50. Solvothermal condensation of HHTP and PBBA in the presence of a

substrate-supported single layer graphene (SLG) surface provides COF-5 as both a

film on the graphene surface, as well as a powder precipitated in the bottom of the

reaction vessel. Reproduced from Ref. [452]. Copyright © 2011, American

Association for the Advancement of Science.

Another recent example of monolayer COFs is a 2D heterotriangulene covalent

polymer. Bieri et al. [454] applied the on-surface synthesis approach to obtain

molecular-thin 2D polymer from tribromo-substituted dimethylmethylene bridged

triphenylamine (DTPA) precursors. The chemical step leading to DTPA films relys on

aryl-aryl homo-coupling between monomers (Fig. 51). By annealing the Ag(111)

surface to temperatures of 400 C, an ultraflat and methyl-cleaved 2D

Page 119: Graphene-analogous low-dimensional materials

118

heterotriangulene polymer with covalent network formed. First-principles studies

[455] demonstrated that the TDPA porous sheet exhibits robust ferromagnetic

half-metallicity under an external strain, which makes the DTPA sheet an ideal

candidate for a spin-selective conductor.

Fig. 51. (a) Structure of DTPA and a fraction of the covalent network. (b) STM

topograph of an ultraflat, methyl-cleaved covalent network obtained by annealing to

400 C, with the structural model and corresponding STM simulation shown in (c).

Reproduced from Ref. [454]. Copyright © 2011, Royal Society of Chemistry.

Compared with the diverse synthesis routes and versatile structures of

coordination polymers, the construction of COFs has been restricted to only a limited

number of monomers, and the lack of suitable procedures that utilized other molecular

units has hindered further advancement of this emerging field. Besides, COFs are

usually hard to characterize since COFs have no single crystal structures. It is thus

important to extend the limited number of synthetic methods and monomer units

Page 120: Graphene-analogous low-dimensional materials

119

available and improve the structural crystallity to go forward these emerging

materials.

7. Conclusion and prospective

The explosive studies on graphene have generated great enthusiasm towards the

explorations of graphene-analogous materials, which is a novel field that has received

growing interests at a spectacular pace in recent years. Different from graphene only

composed of carbon, graphene-analogous materials have more structural, bonding and

functional versatility. The distinctive properties of various graphene-analogous

materials make them promising candidates as 2D alternatives of graphene, and an

impressively large number of papers have appeared in this rising area during the past

few years.

The underlying goal of this review is to document and systematize recent

progress in experimental and theoretical exploitations of graphene-like

low-dimensional materials (2D nanosheets and 1D nanoribbons), including planar

graphene-like materials (h-BN, silicene, and BC3), theoretically predicted planar

materials (SiC, SiC2, B, and B2C), non-planar materials (metal dichalcogenides, metal

oxides and hydroxides, and MXene), metal coordination polymers, and organic

covalent polymers. In our overview, we mainly pay attention to the experimental

synthesis, characterization, functional applications, and theoretical predictions on

stability, electronic and magnetic properties of different systems.

Among them, h-BN is particularly unique due to its structural resemblance to

graphene but with totally distinct properties. BN nanosheets/ribbons have been

Page 121: Graphene-analogous low-dimensional materials

120

synthesized by many techniques, and the structural defects such as vacancies were

observed. Although BN is characterized as a wide-band-gap semiconductor, its

electronic and magnetic properties can be modified to a large scale by defects, doping,

and surface or edge functionalizations. Si has the same valence number as C, yet

single layer Si, silicene, is not planar as graphene bur forms a slightly corrugated

configuration. The Si-C hybrid phase, SiC and SiC2, can form stable planar geometry.

Boron is a fascinating element, and its bonding is featured with an electron-deficient

character. The planar boron-containing materials, including B, BC3, and B2C

nanosheets have been proposed. MoS2 and WS2 nanosheets/ribbons have been

successfully produced, and they have moderate band gaps, suggesting the prospective

applications in nanoelectronics and photoelectronics. 2D nanosheets of layered oxides

and hydroxides can be fabricated via chemical exfoliation, and the resulting

nanosheets always form colloidal suspensions. Particularly, layered V2O5 was

exfoliated into thin nanosheets, which are promising cathode materials for Li-ion

batteries. Moreover, polar (0001) films of wurtzite ZnO with a few atomic layers

would transform into more stable graphitic structures, and the graphene-like ZnO

sheets have been observed in experiment. MAX phase is another family of layered

materials with strongly interacted MX and A atomic layers, and its separated MX

layers (“MXene”) have great potential as anode materials for Li-ion batteries. Besides,

2D coordination polymers and organic covalent polymers with tunable architectures

are another interesting area. They possess unique electromagnetic properties, and can

be additionally applied for gas storage and detection.

Page 122: Graphene-analogous low-dimensional materials

121

Compared with the intensive research focus on graphene, investigations on

graphene-analogous materials have just come into the very beginning, and many

challenges exist, such as their large-scale fabrications and definitive characterizations

with desired thickness. Besides, many properties of graphene-analogous materials are

not fully explored, such as the practically measured mechanical, electronic, magnetic,

and optical properties, and their possible applications are still waiting for the further

explorations. However, we are highly expected that this emerging rich area has a

bright future and will rapidly grow to be an advancing field, and will generate

exciting properties and amazing applications.

Acknowledgements

This work was supported by NSFC (21073096 and 21273118) in China.

References

[1] Novoselov KS, Geim AK, Morozov SV, Jiang D, Zhang Y, Dubonos SV,

Grigorieva IV, Firsov AA. Electric field effect in atomically thin carbon films.

Science 2004;306:666-9.

[2] Geim AK, Novoselov KS. The rise of graphene. Nat Mater 2007;6:183-91.

[3] Singh V, Joung D, Zhai L, Das S, Khondaker SI, Seal S. Graphene based materials:

past, present and future. Prog Mater Sci 2011;56:1178-271.

[ 4 ] Kuila T, Bose S, Mishra AK, Khanra P, Kim NH, Lee JH. Chemical

functionalization of graphene and its applications. Prog Mater Sci 2012;57:

1061-105.

[5] Guo S, Dong S. Graphene nanosheet: synthesis, molecular engineering, thin film,

Page 123: Graphene-analogous low-dimensional materials

122

hybrids, and energy and analytical applications. Chem Soc Rev 2011;40:2644-72.

[6] Tang Q, Zhou, Z, Chen ZF. Graphene-related nanomaterials: Tuning properties by

functionalization. Nanoscale DOI: 10.1039/C3NR33218G

[7] Mas-Ballesté R, Gómez-Navarro C, Gómez-Herrero J, Zamora F. 2D materials: to

graphene and beyond. Nanoscale, 2011;3:20-30.

[8] Xu M, Liang T, Shi M, Chen H. Graphene-like two-dimensional materials. Chem

Rev DOI: 10.1021/cr300263a.

[9] Butler SZ, Hollen SM, Cao L, Cui Y, Gupta JA, Gutiérrez HR, Heinz TF, Hong

SS, Huang J, Ismach AF, Johnston-Halperin E, Kuno M, Plashnitsa VV, Robinson

RD, Ruoff RS, Salahuddin S, Shan J, Shi L, Spencer MG, Terrones M, Windl W,

Goldberger JE. Progress, challenges, and opportunities in two-dimensional

materials beyond graphene. ACS Nano DOI: 10.1021/nn400280c.

[10] Novoselov KS, Jiang D, Schedin F, Booth TJ, Khotkevich VV, Morozov SV,

Geim AK. Two-dimensional atomic crystals. Proc Natl Acad Sci U S A 2005;

102:10451-3.

[11] Ueno A, Fujita T, Matsue M, Yanagisawa H, Oshima C, Patthey F, Ploigt HC,

Schneider WD, Otani S. Scanning tunneling microscopy study on a BC3 covered

NbB2(0001) surface. Surf Sci 2006;600:3518-21.

[12] Kara A, Enriquez H, Seitsonen AP, Lew Yan Voon LC, Vizzini S, Aufray B,

Oughaddou H. A review on silicone- new candidate for electronics. Surf Sci Rep

2012;67:1-18.

[13] Naguib M, Kurtoglu M, Presser V, Lu J, Niu J, Heon M, Hultman L, Gogotsi Y,

Barsoum MW. Two-dimensional nanocrystals produced by exfoliation of

Ti3AlC2. Adv Mater 2011;23:4248-53.

Page 124: Graphene-analogous low-dimensional materials

123

[14] Naguib M, Mashtalir O, Carle J, Presser V, Lu J, Hultman L, Gogotsi Y, Barsoum

MW. Two-dimensional transition metal carbides. ACS Nano 2012;6:1322-31.

[15] Amo-Ochoa P, Welte L, González-Prieto R, Sanz Miguel PJ, Gó mez-García CJ,

Mateo-Martí E, Delgado S, Gómez-Herrero J, Zamora F. Single layers of a

multifunctional laminar Cu(I,II) coordination polymer. Chem Commun 2010;46:

3262-4.

[16] Abel M, Clair S, Ourdjini O, Mossoyan M, Porte L. Single layer of polymeric

Fe-Phthalocyanine: an organometallic sheet on metal and thin insulating film. J

Am Chem Soc 2011;133:1203-5.

[17] Coleman JN, Lotya M, O'Neill A, Bergin SD, et al. Two-dimensional nanosheets

produced by liquid exfoliation of layered materials. Science 2011;331:568-71.

[18] Zeng Z, Yin Z, Huang X, Li H, He Q, Lu G, Boey F, Zhang H. Single-layer

semiconducting nanosheets: high-yield preparation and device fabrication.

Angew Chem Int Ed 2011;50:11093-7.

[19] Zeng H, Zhi C, Zhang Z, Wei X, Wang X, Guo W, Bando Y, Golberg D. “White

graphenes”: boron nitride nanoribbons via boron nitride nanotube unwrapping.

Nano Lett 2010;10:5049-55.

[20] Erickson KJ, Gibb AL, Sinitskii A, Rousseas M, Alem N, Tour JM, Zettl AK.

Longitudinal splitting of boron nitride nanotubes for the facile synthesis of high

quality boron nitride nanoribbons. Nano Lett 2011;11:3221-6.

[21] Wang Z, Li H, Liu Z, Shi Z, Lu J, Suenaga K, Joung S-K, Okazaki T, Gu Z,

Zhou J, Gao Z, Li G, Sanvito S, Wang E, Iijima S. Mixed low-dimensional

nanomaterial: 2D ultranarrow MoS2 inorganic nanoribbons encapsulated in

quasi-1D carbon nanotubes. J Am Chem Soc 2010;132:13840-7.

[22] Wang Z, Zhao K, Li H, Liu Z, Shi Z, Lu J, Suenaga K, Joung SK, Okazaki T, Jin

Page 125: Graphene-analogous low-dimensional materials

124

Z, Gu Z, Gao Z, Iijima S. Ultra-narrow WS2 nanoribbons encapsulated in carbon

nanotubes. J Mater Chem 2011;21:171-80.

[23] Lin Y, Connell JW. Advances in 2D boron nitride nanostructures: nanosheets,

nanoribbons, nanomeshes, and hybrids with graphene. Nanoscale

2012;4:6908-39.

[24] Gao XF, Zhou Z, Zhao YL, Nagase S, Zhang SB, Chen ZF. Comparative study of

carbon and BN nanographenes: ground electronic states and energy gap

engineering. J Phys Chem C 2008;112:12677-82.

[25] Zheng F, Zhou G, Liu Z, Wu J, Duan W, Gu BL, Zhang SB. Half metallicity

along the edge of zigzag boron nitride nanoribbons. Phys Rev B

2008;78:205415.

[26] Chen W, Li Y, Yu G, Li C, Zhang SB, Zhou Z, Chen Z. Hydrogenation: a simple

approach to realize semiconductor-half-metal-metal transition in boron nitride

nanoribbons. J Am Chem Soc 2010;132:1699-705.

[27] Sun L, Li Y, Li Z, Li Q, Zhou Z, Chen Z, Yang J, Hou JG. Electronic structures

of SiC nanoribbons. J Chem Phys 2008;129:174114.

[28] Li Y, Zhou Z, Zhang S, Chen Z. MoS2 nanoribbons: high stability and unusual

electronic and magnetic properties. J Am Chem Soc 2008;130:16739-44.

[29] Li Y, Li F, Zhou Z, Chen Z. SiC2 silagraphene and its one-dimensional

derivatives: where planar tetracoordinate silicon happens. J Am Chem Soc

2011;133:900-8.

[30] Wang QH, Kalantar-Zadeh K, Kis A, Coleman JN, Strano MS. Electronics and

optoelectronics of two-dimensional transition metal dichalcogenides. Nat

Nanotech 2012;7:699-712.

Page 126: Graphene-analogous low-dimensional materials

125

[31] Pacilé D, Meyer JC, Girit CO, Zettl A. The two-dimensional phase of boron

nitride: few-atomic-layer sheets and suspended membranes. Appl Phys Lett

2008;92:133107.

[32] Gorbachev RV, Riaz I, Nair RR, Jalil R, Britnell L, Belle BD, Hill EW,

Novoselov KS, Watanabe K, Taniguchi T, Geim AK, Blake P. Hunting for

monolayer boron nitride: optical and raman signatures. Small 2011;7:465-8.

[33] Lee C, Yan H, Brus LE, Heinz TF, Hone J, Ryu S. Anomalous lattice vibrations

of single and few-layer MoS2. ACS Nano 2010;4:2695-700.

[34] Hu P, Wen Z, Wang L, Tan P, Xiao K. Synthesis of few-Layer GaSe nanosheets

for high performance photodetectors. ACS Nano 2012;6:5988-94.

[35] Late DJ, Liu B, Luo J, Yan A, Ramakrishna Matte HSS, Grayson M, Rao CNR,

Dravid VP. GaS and GaSe ultrathin layer transistors. Adv Mater

2012;24:3549-54.

[36] Hossain MZ, Rumyantsev SL, Shahil KMF, Teweldebrhan D, Shur M, Balandin

AA. Low-frequency current fluctuations in “graphene-like” exfoliated thin-films

of bismuth selenide topological insulators. ACS Nano 2011;5:2657-63.

[37] Teweldebrhan D, Goyal V, Balandin AA. Exfoliation and characterization of

bismuth telluride atomic quintuples and quasi-two-dimensional crystals. Nano

Lett 2010;10:1209-18.

[38] Han W, Wu L, Zhu Y, Watanabe K, Taniguchi T. Structure of chemically

derived mono- and few-atomic layer boron nitride sheets. Appl Phys Lett 2008;

93:223103.

[39] Zhi C, Bando Y, Tang C, Kuwahara H, Golberg D. Largescale fabrication of

boron nitride nanosheets and their utilization in polymeric composites with

Page 127: Graphene-analogous low-dimensional materials

126

improved thermal and mechanical properties. Adv Mater 2009;21:2889-93.

[40] Lin Y, Williams TV, Xu TB, Cao W, Elsayed-Ali HE, Connell JW. Aqueous

dispersions of few-layered and monolayered hexagonal boron nitride nanosheets

from sonication-assisted hydrolysis: critical role of water. J Phys Chem C

2011;115:2679-85.

[41] Kim JS, Borisenko KB, Nicolosi V, Kirkland AI. Controlled radiation damage

and edge structures in boron nitride membranes. ACS Nano 2011;5:3977-86.

[42] Wang Y, Shi Z, Yin J. Boron nitride nanosheets: large-scale exfoliation in

methanesulfonic acid and their composites with polybenzimidazole. J Mater

Chem 2011;21:11371-7.

[43] Warner JH, Rümmeli MH, Bachmatiuk A, Büchner B. Atomic resolution

imaging and topography of boron nitride sheets produced by chemical

exfoliation. ACS Nano 2010;4:1299-304.

[44] Cunningham G, Lotya M, Cucinotta CS, Sanvito S, Bergin SD, Menzel R,

Shaffer MSP, Coleman JN. Solvent exfoliation of transition metal

dichalcogenides: dispersibility of exfoliated nanosheets varies only weakly

between compounds. ACS Nano 2012;6:3468-80.

[45] Zhou KG, Mao NN, Wang HX, Peng Y, Zhang HL. A mixed-solvent strategy for

efficient exfoliation of inorganic graphene analogues. Angew Chem Int Ed

2011;50:10839-42.

[46] Smith RJ, King PJ, Lotya M, Wirtz C, Khan U, De S, O’Neill A, Duesberg GS,

Grunlan JC, Moriarty G, Chen J, Wang J, Minett AI, Nicolosi V, Coleman JN.

Large-scale exfoliation of inorganic layered compounds in aqueous surfactant

solutions. Adv Mater 2011;23:3944-8.

Page 128: Graphene-analogous low-dimensional materials

127

[47] Feng J, Peng L, Wu C, Sun X, Hu S, Lin C, Dai J, Yang J, Xie Y. Giant moisture

responsiveness of VS2 ultrathin nanosheets for novel touchless positioning

interface. Adv Mater 2012;24:1969-74.

[48] Rui X, Lu Z, Yu H, Yang D, Hng HH, Lim TM, Yan Q. Ultrathin V2O5

nanosheet cathodes: realizing ultrafast reversible lithium storage. Nanoscale

2013;5:556-60.

[49] Li X, Hao X, Zhao M, Wu Y, Yang J, Tian Y, Qian G. Exfoliation of hexagonal

boron nitride by molten hydroxides. Adv Mater 2013;25:2200-4.

[50] O’Neill A, Khan U, Coleman JN. Preparation of high concentration dispersions

of exfoliated MoS2 with increased flake size. Chem Mater 2012;24:2414-21.

[51] Ramakrishna Matte HSS, Plowman B, Datta R, Rao CNR. Graphene analogues

of layered metal selenides. Dalton Trans 2011;40:10322-5.

[52] Lee K, Kim HY, Lotya M, Coleman JN, Kim GT, Duesberg GS. Electrical

characteristics of molybdenum disulfide flakes produced by liquid exfoliation.

Adv Mater 2011;23:4178-82.

[53] Lin Y, Williams TV, Connell JW. Soluble, exfoliated hexagonal boron nitride

nanosheets. J Phys Chem Lett 2010;1:277-83.

[ 54 ] Lin Y, Williams TV, Cao W, Elsayed-Ali HE, Connell JW. Defect

functionalization of hexagonal boron nitride nanosheets. J Phys Chem C 2010;

114:17434-9.

[55] Nazarov AS, Demin VN, Grayfer ED, Bulavchenko AI, Arymbaeva AT, Shin HJ,

Choi JY, Fedorov VE. Functionalization and dispersion of hexagonal boron

nitride (h-BN) nanosheets treated with inorganic reagents. Chem Asian J 2012;

7:554-60.

Page 129: Graphene-analogous low-dimensional materials

128

[56] Sainsbury T, Satti A, May P, O’Neill A, Nicolosi V, Gun’ko YK, Coleman JN.

Covalently functionalized hexagonal boron nitride nanosheets by nitrene addition.

Chem Eur J 2012;18:10808-12.

[57] Osada M, Sasaki T. Two-dimensional dielectric nanosheets: novel

nanoelectronics from nanocrystal building blocks. Adv Mater 2012;24:210-28.

[58] Ma R, Sasaki T. Nanosheets of oxides and hydroxides: ultimate 2D

charge-bearing functional crystallites. Adv Mater 2010;22:5082-104.

[59] Gordon RA, Yang D, Crozier ED, Jiang DT, Frindt RF. Structures of exfoliated

single layers of WS2 , MoS2 , and MoSe2 in aqueous suspension. Phys Rev B

2002;65:125407.

[60] Zeng Z, Sun T, Zhu J, Huang X, Yin Z, Lu G, Fan Z, Yan Q, Hng HH, Zhang H.

An effective method for the fabrication of few-layer-thick inorganic nanosheets.

Angew Chem Int Ed 2012;51:9052-6.

[61] Eda G, Fujita T, Yamaguchi H, Voiry D, Chen M, Chhowalla M. Coherent

atomic and electronic heterostructures of single-layer MoS2. ACS Nano 2012;

6:7311-7.

[62] Corso M, Auwärter W, Muntwiler M, Tamai A, Greber T, Osterwalder J. Boron

nitride nanomesh. Science 2004;303:217-20.

[63] Sutter P, Lahiri J, Albrecht P, Sutter E. Chemical vapor deposition and etching of

high-quality monolayer hexagonal boron nitride films. ACS Nano 2011;5:

7303-9.

[64] Joshi S, Ecija D, Koitz R, Iannuzzi M, Seitsonen AP, Hutter J, Sachdev H,

Vijayaraghavan S, Bischoff F, Seufert K, Barth JV, Auwarter W. Boron nitride

on Cu(111): an electronically corrugated monolayer. Nano Lett 2012;12:5821-8.

Page 130: Graphene-analogous low-dimensional materials

129

[65] Berner S, Corso M, Widmer R, Groening O, Laskowski R, Blaha P, Schwarz K,

Goriachko A, Over H, Gsell S, Schreck M, Sachdev H, Greber T, Osterwalder J.

Boron nitride nanomesh: functionality from a corrugated monolayer. Angew

Chem Int Ed 2007;46:5115-9.

[66] Shi Y, Hamsen C, Jia X, Kim KK, Reina A, Hofmann M, Hsu AL, Zhang K, Li

H, Juang ZY, Dresselhaus MS, Li LJ, Kong J. Synthesis of few-layer hexagonal

boron nitride thin film by chemical vapor deposition. Nano Lett 2010;10:4134-9.

[67] Song L, Ci L, Lu H, Sorokin PB, Jin C, Ni J, Kvashnin AG, Kvashnin DG, Lou J,

Yakobson BI, Ajayan PM. Large scale growth and characterization of atomic

hexagonal boron nitride layers. Nano Lett 2010;10:3209-15.

[68] Whittell GR, Manners I. Advances with ammonia-borane: improved recycling

and use as a precursor to atomically thin BN films. Angew Chem Int Ed

2011;50:10288-9.

[ 69 ] Kim KK, Hsu A, Jia X, Kim SM, Shi Y, Hofmann M, Nezich D,

Rodriguez-Nieva JF, Dresselhaus M, Palacios T, Kong J. Synthesis of monolayer

hexagonal boron nitride on Cu foil using chemical vapor deposition. Nano Lett

2012;12:161-6.

[70] Lee KH, Shin HJ, Lee J, Lee I, Kim GH, Choi JY, Kim SW. Large-scale

synthesis of high-quality hexagonal boron nitride nanosheets for large-area

graphene electronics. Nano Lett 2012;12:714-8.

[71] Chatterjee S, Luo Z, Acerce M, Yates DM, Charlie Johnson AT, Sneddon LG.

Chemical vapor deposition of boron nitride nanosheets on metallic substrates via

decaborane/ammonia reactions. Chem Mater 2011;23:4414-6.

[72] Ismach A, Chou H, Ferrer DA, Wu Y, McDonnell S, Floresca HC, Covacevich A,

Page 131: Graphene-analogous low-dimensional materials

130

Pope C, Piner R, Kim MJ, Wallace RM, Colombo L, Ruoff RS. Toward the

controlled synthesis of hexagonal boron nitride films. ACS Nano

2012;6:6378-85.

[73] Pakdel A, Zhi C, Bando Y, Nakayama T, Golberg D. Boron nitride nanosheet

coatings with controllable water repellency. ACS Nano 2011;5:6507-15.

[74] Yu J, Qin L, Hao Y, Kuang S, Bai X, Chong Y, Zhang W, Wang E. Vertically

aligned boron nitride nanosheets: chemical vapor synthesis, ultraviolet light

emission, and superhydrophobicity. ACS Nano 2010;4:414-22.

[75] Liu Z, Song L, Zhao S, Huang J, Ma L, Zhang J, Lou J, Ajayan PM. Direct

growth of graphene/hexagonal boron nitride stacked layers. Nano Lett 2011;

11:2032-7.

[76] Ci L, Song L, Jin C, Jariwala D, Wu D, Li Y, Srivastava A, Wang ZF, Storr K,

Balicas L, Liu F, Ajayan PM. Atomic layers of hybridized boron nitride and

graphene domains. Nat Mater 2010;9:430-5.

[77] Zhan Y, Liu Z, Najmaei S, Ajayan PM, Lou J. Large-area vapor-phase growth

and characterization of MoS2 atomic layers on a SiO2 substrate. Small 2012;8:

966-71.

[78] Lee YH, Zhang XQ, Zhang W, Chang MT, Lin CT, Chang KD, Yu YC, Wang

JTW, Chang CS, Li LJ, Lin TW. Synthesis of large-area MoS2 atomic layers

with chemical vapor deposition. Adv Mater 2012;24:2320-5.

[79] Shi Y, Zhou W, Lu A, Fang W, Lee Y, Hsu AL, Kim SM, Kim KK, Yang HY,

Li L, Idrobo J, Kong J. van der waals epitaxy of MoS2 layers using graphene as

growth templates. Nano Lett 2012;12:2784-91.

[80] Zhang C, Hao X, Wu Y, Du M. Synthesis of vertically aligned boron nitride

Page 132: Graphene-analogous low-dimensional materials

131

nanosheets using CVD method. Mater Res Bull 2012;47:2277-81.

[81] Gao R, Yin L, Wang C, Qi Y, Lun N, Zhang L, Liu Y, Kang L, Wang X.

High-yield synthesis of boron nitride nanosheets with strong ultraviolet

cathodoluminescence emission. J Phys Chem C 2009;113:15160-5.

[82] Sachdev H, Muller F, Hufner, S. Formation of boron-based films and boron

nitride layers by CVD of a boron ester. Angew Chem Int Ed 2011;50:3701-5.

[ 83 ] 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.

[84] Aufray B, Kara A, Vizzini S, Oughaddou H, Léandri C, Ealet B, Le Lay G.

Graphene-like silicon nanoribbons on Ag(110): a possible formation of silicene.

Appl Phys Lett 2010;96:183102.

[85] De Padova P, Quaresima C, Ottaviani C, Sheverdyaeva PM, Moras P, Carbone C,

Topwal D, Olivieri B, Kara A, Oughaddou H, Aufray B, Le Lay G. Evidence of

graphene-like electronic signature in silicene nanoribbons. Appl Phys Lett

2010;96:261905.

[86] Padova PD, Kubo O, Olivieri B, Quaresima C, Nakayama T, Aono M, Lay GL.

Multilayer silicene nanoribbons. Nano Lett 2012;12:5500-3.

[87] Lalmi B, Oughaddou H, Enriquez H, Kara A, Vizzini S, Ealet B, Aufray B.

Epitaxial growth of a silicene sheet. Appl Phys Lett 2010;97:223109.

[88] Lay GL, De Padova P, Resta A, Bruhn T, Vogt P. Epitaxial silicene: can it be

strongly strained? J Phys D Appl Phys 2012;45:392001.

[89] Vogt P, Padova PD, Quaresima C, Avila J, Frantzeskakis E, Asensio MC, Resta

A, Ealet B, Lay GL. Silicene: compelling experimental evidence for

Page 133: Graphene-analogous low-dimensional materials

132

graphenelike two-dimensional silicon. Phys Rev Lett 2012;108:155501.

[90] Chiappe D, Grazianetti C, Tallarida G, Fanciulli M, Molle A. Local electronic

properties of corrugated silicene phases. Adv Mater 2012;24:5088-93.

[91] Feng B, Ding Z, Meng S, Yao Y, He X, Cheng P, Chen L, Wu K. Evidence of

silicene in honeycomb structures of silicon on Ag(111). Nano Lett 2012;12:

3507-11.

[92] Lin CL, Arafune R, Kawahara K, Tsukahara N, Minamitani E, Kim Y, Takagi N,

Kawai M. Structure of silicene grown on Ag(111). Appl Phys Express

2012;5:045802.

[93] Jamgotchian H, Colignon Y, Hamzaoui N, Ealet B, Hoarau J Y, Aufray B,

Biberian J P. Growth of silicene layers on Ag(111): unexpected effect of the

substrate temperature. J Phys Condens Matter 2012;24:172001.

[94] Meng L, Wang Y, Zhang L, Du S, Wu R, Li L, Zhang Y, Li G, Zhou H, Hofer

WA, Gao HJ. Buckled Silicene Formation on Ir(111). Nano Lett

2013;13:685-90.

[95] Fleurence A, Friedlein R, Ozaki T, Kawai H, Wang Y, Yamada-Takamura Y.

Experimental evidence for epitaxial silicene on diboride thin films. Phys Rev

Lett 2012;108:245501.

[96] Li L, Li LH, Chen Y, Dai XJ, Lamb PR, Cheng BM, Lin MY, Liu X.

High-quality boron nitride nanoribbons: unzipping during nanotube synthesis.

Angew Chem Int Ed 2013;52:4212-6.

[97] Alem N, Erni R, Kisielowski C, Rossell MD, Gannett W, Zettl A. Atomically

thin hexagonal boron nitride probed by ultrahigh-resolution transmission electron

microscopy. Phys Rev B 2009;80:155425.

Page 134: Graphene-analogous low-dimensional materials

133

[98] Xu M, Fujita D, Chen H, Hanagata N. Formation of monolayer and few-layer

hexagonal boron nitride nanosheets via surface segregation. Nanoscale

2011;3:2854-8.

[99] Jin C, Lin F, Suenaga K, Iijima S. Fabrication of a freestanding boron nitride

single layer and its defect assignments. Phys Rev Lett 2009;102:195505.

[100] Meyer JC, Chuvilin A, Algara-Siller G, Biskupek J, Kaiser U. Selective

sputtering and atomic resolution imaging of atomically thin boron nitride

membranes. Nano Lett 2009;9:2683-9.

[101] Wang X, Zhi C, Li L, Zeng H, Li C, Mitome M, Golberg D, Bando Y.

“Chemical blowing” of thin-walled bubbles: high-throughput fabrication of

large-area, few-layered BN and Cx-BN nanosheets. Adv Mater

2011;23:4072-6.

[102] Castellanos-Gomez A, Barkelid M, Goossens AM, Calado VE, van der Zant

HSJ, Steele GA. Laser-thinning of MoS2: on demand generation of a

single-layer semiconductor. Nano Lett 2012;12:3187-92.

[103] Liu KK, Zhang W, Lee YH, Lin YC, Chang MT, Su CY, Chang CS, Li H, Shi

Y, Zhang H, Lai CS, Li LJ. Growth of large-area and highly crystalline MoS2

thin layers on insulating substrates. Nano Lett 2012;12:1538-44.

[104] Balendhran S, Ou JZ, Bhaskaran M, Sriram S, Ippolito S, Vasic Z, Kats E,

Bhargava S, Zhuiykov S, Kalantar-zadeh K. Atomically thin layers of MoS2 via

a two step thermal evaporation–exfoliation method. Nanoscale 2012;4:461-6.

[105] Ramakrishna Matte HSS, Gomathi A, Manna AK, Late DJ, Datta R, Pati SK,

Rao CNR. MoS2 and WS2 analogues of graphene. Angew Chem Int Ed

2010;49:4059-62.

Page 135: Graphene-analogous low-dimensional materials

134

[106] Feng J, Sun X, Wu C, Peng L, Lin C, Hu S, Yang J, Xie Y. Metallic

few-layered VS2 ultrathin nanosheets: high two-dimensional conductivity for

in-plane supercapacitors. J Am Chem Soc 2011;133:17832-8.

[107] Plashnitsa VV, Vietmeyer F, Petchsang N, Tongying P, Kosel TH, Kuno M.

Synthetic strategy and structural and optical characterization of thin highly

crystalline titanium disulfide nanosheets. J Phys Chem Lett 2012;3:1554-8.

[108] Taskin AA, Sasaki S, Segawa K, Ando Y. Manifestation of topological

protection in transport properties of epitaxial Bi2Se3 thin films. Phys Rev Lett

2012;109:066803.

[109] Zhao Y, Hughes RW, Su Z, Zhou W, Gregory DH. One-step synthesis of

bismuth telluride nanosheets of a few quintuple layers in thickness. Angew

Chem Int Ed 2011;50:10397-401.

[110] Nag A, Raidongia K, Hembram KPSS, Datta R, Waghmare UV, Rao CNR.

Graphene analogues of BN: novel synthesis and properties. ACS Nano 2010;

4:1539-44.

[111] Raidongia K, Nag A, Hembram KPSS, Waghmare UV, Datta R, Rao CNR.

BCN: a graphene analogue with remarkable adsorptive properties. Chem Eur J

2010;16:149-57.

[112] Wang X, Pakdel A, Zhi C, Watanabe K, Sekiguchi T, Golberg D, Bando Y.

High-yield boron nitride nanosheets from 'chemical blowing': towards practical

applications in polymer composites. J Phys Condens Matter 2012;24:314205.

[113] Wu Z, Wang D, Wang Y, Sun A. Preparation and tribological properties of

MoS2 nanosheets. Adv Eng Mater 2012;12:534-8.

[114] Liang Y, Feng R, Yang S, Ma H, Liang J, Chen J. Rechargeable Mg batteries

Page 136: Graphene-analogous low-dimensional materials

135

with graphene-like MoS2 cathode and ultrasmall Mg nanoparticle anode. Adv

Mater 2011;23:640-3.

[115] Seo JW, Jun YW, Park SW, Nah H, Moon T, Park B, Kim JG, Kim YJ, Cheon J.

Two-dimensional nanosheet crystals. Angew Chem Int Ed 2007;46:8828-31.

[116] Li H, Lu G, Yin Z, He Q, Li H, Zhang Q, Zhang H. Optical identification of

single- and few-layer MoS2 sheets. Small 2012;8:682-6.

[117] Perdew JP, Wang Y. Accurate and simple density functional for the electronic

exchange energy: generalized gradient approximation. Phys Rev B

1986;33:8800-2.

[118] Perdew JP, Wang Y. Accnrate and simple analytic representation of the

electron-gas correlation energy. Phys Rev B 1992;45:13244-9.

[119] Perdew JP, Burke K, Ernzerhof M. Generalized gradient approximation made

simple. Phys Rev Lett 1996;77:3865-8.

[120] Heyd J, Scuseria GE, Ernzerhof M. Hybrid functionals based on a screened

Coulomb potential. J Chem Phys 2003;118:8207-15.

[121] Delley B. An all-electron numerical-method for solving the local density

functional for polyatomic-molecules. J Chem Phys 1990;92:508-17.

[122] Delley B. From molecules to solids with the DMol(3) approach. J Chem Phys

2000;113:7756-64.

[123] Ordejon P, Artacho E, Soler JM. Selfconsistent order-N density-functional

calculations for very large systems. Phys Rev B 1996;53:R10441.

[124] Soler JM, Artacho E, Gale JD, Garcia A, Junquera J, Ordejon P, Sánchez-Portal

D. The SIESTA method for ab initio order-N materials simulation. J Phys

Page 137: Graphene-analogous low-dimensional materials

136

Condens Matter 2002;14:2745-79.

[125] Kresse G, Furthmullerr J. Efficient iterative schemes for ab initio total-energy

calculations using a plane-wave basis set. Phys Rev B 1996;54:11169-86.

[126] Blöchl PE. Projector augmented-wave method. Phys Rev B 1994;50:17953-79.

[ 127 ] Dolg M, Wedig U, Stoll H, Preuss H. Energy-adjusted ab initio

pseudopotentials for the first row transition elements. J Chem Phys

1987;86:866.

[128] Delley B. Hardness conserving semilocal pseudopotentials. Phys Rev B

2002;66:155125.

[129] Troullier N, Martins JL. Efficient pseudopotentials for plane-wave calculations.

Phys Rev B 1991;43:1993-2006.

[130] Vanderbilt D. Soft self-consistent pseudopotentials in a generalized eigenvalue

formalism. Phys Rev B 1990;41:7892-5.

[131] Segall MD, Lindan PJD, Probert MJ, Pickard CJ, Hasnip PJ, Clark SJ, Payne

MC. First-principles simulation: ideas, illustrations and the CASTEP code. J

Phys Condens Matter 2002;14:2717-44.

[132] Giannozzi P, Baroni S, et al. QUANTUM ESPRESSO: a modular and

open-source software project for quantum simulations of materials. J Phys

Condens Matter 2009;21:395502.

[133] Gonze X, Beuken JM, Caracas R, Detraux F, et al. First-principles computation

of material properties: the ABINIT software project. Comput Mater Sci

2002;25:478-92.

[134] Friedrich C, Betzinger M, Schlipf M, Blügel S, Schindlmayr A. Hybrid

functionals and GW approximation in the FLAPW method. J Phys Condens

Page 138: Graphene-analogous low-dimensional materials

137

Matter 2012;24:293201.

[135] Cococcioni M, de Gironcoli S. Linear response approach to the calculation of

the effective interaction parameters in the LDA+U method. Phys Rev B

2005;71:035105.

[136] Gräfenstein J, Cremer D. An efficient algorithm for the density-functional

theory treatment of dispersion interactions. J Chem Phys 2009;130:124105.

[137] Blase X, Rubio A, Louie SG, Cohen ML. Quasiparticle band structure of bulk

hexagonal boron nitride and related systems. Phys Rev B 1995;51:6868-75.

[138] Watanabe K, Taniguchi T, Kanda H. Direct-bandgap properties and evidence

for ultraviolet lasing of hexagonal boron nitride single crystal. Nat Mater

2004;3:404-9.

[ 139 ] Topsakal M, Aktürk E, Ciraci S. First-principles study of two- and

one-dimensional honeycomb structures of boron nitride. Phys Rev B

2009;79:115442.

[140] Barone V, Peralta JE. Magnetic boron nitride nanoribbons with tunable

electronic properties. Nano Lett 2008;8:2210-4.

[141] Hu C, Ogura R, Onoda N, Konabe S, Watanabe K. Quasiparticle band gaps of

boron nitride nanoribbons. Phys Rev B 2012;85:245420.

[142] Mukherjee R, Bhowmick S. Edge stabilities of hexagonal boron nitride

nanoribbons: a first-principles study. J Chem Theory Comput 2011;7:720-4.

[143] Kan E, Wu F, Xiang H, Yang J, Whangbo MH. Half-metallic Dirac point in

B-edge hydrogenated BN nanoribbons. J Phys Chem C 2011;115:17252-4.

[144] Lai L, Lu J, Wang L, Luo G, Zhou J, Qin R, Gao Z, Mei WN. Magnetic

properties of fully bare and half-bare boron nitride nanoribbons. J Phys Chem C

Page 139: Graphene-analogous low-dimensional materials

138

2009;113:2273-6.

[145] Park CH, Louie SG. Energy gaps and stark effect in boron nitride nanoribbons. Nano Lett 2008;8:2200-3.

[146] Ding Y, Wang Y, Ni J. The stabilities of boron nitride nanoribbons with

different hydrogen-terminated edges. Appl Phys Lett 2009;94:233107.

[147] Suenaga K, Kobayashi H, Koshino M. Core-level spectroscopy of point defects

in single layer h-BN. Phys Rev Lett 2012;108:075501.

[148] Yin LC, Cheng HM, Saito R. Triangle defect states of hexagonal boron nitride

atomic layer: density functional theory calculations. Phys Rev B

2010;81:153407.

[149] Du A, Chen Y, Zhu Z, Amal R, Lu GQ, Smith SC. Dots versus antidots:

computational exploration of structure, magnetism, and half-metallicity in

boron-nitride nanostructures. J Am Chem Soc 2009;131:17354-9.

[150] Okada S. Atomic configurations and energetics of vacancies in hexagonal boron

nitride: first-principles total-energy calculations. Phys Rev B 2009;80:161404

(R).

[151] Chen W, Li Y, Yu G, Zhou Z, Chen Z. Electronic structure and reactivity of

boron nitride nanoribbons with Stone-Wales defects. J Chem Theory Comput

2009;5:3088-95.

[ 152 ] Liu Y, Zou X, Yakobson BI. Dislocations and grain boundaries in

two-dimensional boron nitride. ACS Nano 2012;6:7053-8.

[153] Pan Y, Yang Z. Electronic structures and spin gapless semiconductors in BN

nanoribbons with vacancies. Phys Rev B 2010;82:195308.

[154] Auwärter W, Muntwiler M, Osterwalder J, Greber T. Defect lines and

Page 140: Graphene-analogous low-dimensional materials

139

two-domain structure of hexagonal boron nitride films on Ni(1 1 1). Surf Sci

2003;545:L735-40.

[155] Li X, Wu X, Zeng XC, Yang J. Band-gap engineering via tailored line defects

in boron-nitride nanoribbons, sheets, and nanotubes. ACS Nano

2012;6:4104-12.

[156] Ataca C, Ciraci S. Functionalization of BN honeycomb structure by adsorption

and substitution of foreign atoms. Phys Rev B 2010;82:165402.

[157] Huang B, Xiang H, Yu J, Wei SH. Effective control of the charge and magnetic

states of transition-metal atoms on single-layer boron nitride. Phys Rev Lett

2012;108:206802.

[158] Zhou YG, Yang P, Sun X, Wang ZG, Zu XT, Gao F. First-principles study of

the noble metal-doped BN layer. J Appl Phys 2011;109:084308.

[159] Liu RF, Cheng C. Ab initio studies of possible magnetism in a BN sheet by

nonmagnetic impurities and vacancies. Phys Rev B 2007;76:014405.

[160] Du AJ, Smith SC, Lu GQ. First-principle studies of electronic structure and

C-doping effect in boron nitride nanoribbons. Chem Phys Lett 2007;447:181-6.

[161] Wei X, Wang MS, Bando Y, Golberg D. Electron-beam-induced substitutional

carbon doping of boron nitride nanosheets, nanoribbons, and nanotubes. ACS

Nano 2011;5:2916-22.

[162] Krivanek OL, Chisholm MF, Nicolosi V, Pennycook TJ, Corbin GJ, Dellby N,

Murfitt MF, Own CS, Szilagyi ZS, Oxley MP, Pantelides ST, Pennycook SJ.

Atom-by-atom structural and chemical analysis by annular dark-field electron

microscopy. Nature 2010;464:571-4.

[163] Berseneva N, Krasheninnikov AV, Nieminen RM. Mechanisms of postsynthesis

Page 141: Graphene-analogous low-dimensional materials

140

doping of boron nitride nanostructures with carbon from first-principles

simulations. Phys Rev Lett 2011;107:035501.

[164] Tang Q, Zhou Z, Chen Z. Molecular charge transfer: a simple and effective

route to engineer the band structures of BN nanosheets and nanoribbons. J Phys

Chem C 2011;115:18531-7.

[165] Tang Q, Zhou Z, Shen P, Chen Z. Band gap engineering of BN sheets by

interlayer dihydrogen bonding and electric field control. ChemPhysChem DOI:

10.1002/cphc.201300141.

[ 166 ] Bhattacharya A, Bhattacharya S, Das GP. Band gap engineering by

functionalization of BN sheet. Phys Rev B 2012;85:035415.

[167] Zhou J, Wang Q, Sun Q, Jena P. Electronic and magnetic properties of a BN

sheet decorated with hydrogen and fluorine. Phys Rev B 2010;81:085442.

[168] Ma Y, Dai Y, Guo M, Niu C, Yu L, Huang B. Strain-induced magnetic

transitions in half-fluorinated single layers of BN, GaN and graphene.

Nanoscale 2011;3:2301-6.

[169] Zhang Z, Zeng XC, Guo W. Fluorinating hexagonal boron nitride into

diamond-like nanofilms with tunable band gap and ferromagnetism. J Am

Chem Soc 2011;133:14831-8.

[170] Zhao Y, Wu X, Yang J, Zeng XC. Oxidation of a two-dimensional hexagonal

boron nitride monolayer: a first-principles study. Phys Chem Chem Phys

2012;14:5545-50.

[171] Zhang HX, Feng PX. Controlling bandgap of rippled hexagonal boron nitride

membranes via plasma treatment. ACS Appl Mater Interfaces 2012;4:30-3.

[172] Wu X, Wu M, Zeng XC. Chemically decorated boron-nitride nanoribbons.

Page 142: Graphene-analogous low-dimensional materials

141

Front Phys China 2009;4:367-72.

[ 173 ] Lopez-Bezanilla A, Huang J, Terrones H, Sumpter BG. Boron nitride

nanoribbons become metallic. Nano Lett 2011;11:3267-73.

[174] Lopez-Bezanilla A, Huang J, Terrones H, Sumpter BG. Structure and electronic

properties of edge-functionalized armchair boron nitride nanoribbons. J Phys

Chem C 2012;116:15675-81.

[175] Wang Y, Ding Y, Ni J. Electronic structures of Fe-terminated armchair boron

nitride nanoribbons. Appl Phys Lett 2011;99:053123.

[176] Giovannetti G, Khomyakov PA, Brocks G, Kelly PJ, van den Brink J.

Substrate-induced band gap in graphene on hexagonal boron nitride: ab initio

density functional calculations. Phys Rev B 2007;76:073103.

[ 177 ] Kaloni TP, Cheng YC, Schwingenschlögl U. Electronic structure of

superlattices of graphene and hexagonal boron nitride. J Mater Chem

2012;22:919-22.

[178] Sakai Y, Koretsune T, Saito S. Electronic structure and stability of layered

superlattice composed of graphene and boron nitride monolayer. Phys Rev B

2011;83:205434.

[179] Zhong X, Amorim RG, Scheicher RH, Pandey R, Karna SP. Electronic

structure and quantum transport properties of trilayers formed from graphene

and boron nitride. Nanoscale 2012;4:5490-8.

[180] Ramasubramaniam A, Naveh D, Towe E. Tunable band gaps in bilayer

graphene-BN heterostructures. Nano Lett 2011;11:1070-5.

[181] Li Y, Sun Q, Chen L, Zhou P, Wang P, Ding S, Zhang DW. Hexagonal boron

nitride intercalated multi-layer graphene: a possible ultimate solution to

Page 143: Graphene-analogous low-dimensional materials

142

ultra-scaled interconnect technology. AIP Adv 2012;2:012191.

[182] Kan E, Ren H, Wu F, Li Z, Lu R, Xiao C, Deng K, Yang J. Why the band gap

of graphene is tunable on hexagonal boron nitride. J Phys Chem C

2012;116:3142-6.

[183] Kharche N, Nayak SK. Quasiparticle band gap engineering of graphene and

graphone on hexagonal boron nitride substrate. Nano Lett 2011;11:5274-8.

[184] Zhong X, Yap YK, Pandey R, Karna SP. First-principles study of strain-induced

modulation of energy gaps of graphene/BN and BN bilayers. Phys Rev B

2011;83:193403.

[185] Bokdam M, Khomyakov PA, Brocks G, Zhong Z, Kelly PJ. Electrostatic

doping of graphene through ultrathin hexagonal boron nitride films. Nano Lett

2011;11:4631-5.

[186] Xu Y, Guo Z, Chen H, Yuan Y, Lou J, Lin X, Gao H, Chen H, Yu B. In-plane

and tunneling pressure sensors based on graphene/hexagonal boron nitride

heterostructures. Appl Phys Lett 2011;99:133109.

[187] Ding X, Ding G, Xie X, Huang F, Jiang M. Direct growth of few layer graphene

on hexagonal boron nitride by chemical vapor deposition. Carbon

2011;49:2522-5.

[188] Gannett W, Regan W, Watanabe K, Taniguchi T, Crommie MF, Zetti A. Boron

nitride substrate for high mobility chemical vapor deposited graphene. Appl

Phys Lett 2011;98:242105.

[189] Kim E, Yu T, Song ES, Yu B. Chemical vapor deposition-assembled graphene

field-effect transistor on hexagonal boron nitride. Appl Phys Lett 2011;98:

262103.

Page 144: Graphene-analogous low-dimensional materials

143

[190] Bjelkevig C, Mi Z, Xiao J, Dowben PA, Wang L, Mei WN, Kelber JA.

Electronic structure of a graphene/hexagonal-BN heterostructure grown on Ru

(0001) by chemical vapor deposition and atomic layer deposition: extrinsically

doped graphene. J Phys Condens Matter 2010;22:302002.

[191] Usachov D, Adamchuk VK, Haberer D, Grüneis A, Sachdev H, Preobrajenski

AB, Laubschat C, Vyalikh DV. Quasifreestanding single-layer hexagonal boron

nitride as a substrate for graphene synthesis. Phys Rev B 2010;82:075415.

[192] Bresnehan MS, Hollander MJ, Wetherington M, LaBella M, Trumbull KA,

Cavalero R, Snyder DW, Robinson JA. Integration of hexagonal boron nitride

with quasi-freestanding epitaxial graphene: toward wafer-scale,

high-performance devices. ACS Nano 2012;6:5234-41.

[193] Son M, Lim H, Hong M, Choi HC. Direct growth of graphene pad on exfoliated

hexagonal boron nitride surface. Nanoscale 2011;3:3089–93.

[194] Lin T, Wang Y, Bi H, Wan D, Huang F, Xie X, Jiang M. Hydrogen flame

synthesis of few-layer graphene from a solid carbon source on hexagonal boron

nitride. J Mater Chem 2012;22:2859-62.

[195] Haigh SJ, Gholinia A, Jalil R, Romani S, Britnell L, Elias DC, Novoselov KS,

Ponomarenko LA, Geim AK, Gorbachev R. Cross-sectional imaging of

individual layers and buried interfaces of graphene-based heterostructures and

superlattices. Nat Mater 2012;11:764-7.

[196] Garcia AGF, Neumann M, Amet F, Williams JR, Watanabe K, Taniguchi T,

Goldhaber-Gordon D. Effective cleaning of hexagonal boron nitride for

graphene devices. Nano Lett 2012;12:4449-54.

[197] Meric I, Dean C, Young A, Hone J, Kim P, Shepard KL. Graphene field-effect

Page 145: Graphene-analogous low-dimensional materials

144

transistors based on boron nitride gate dielectrics. IEDM Tech Dig 2010;10:

556-9.

[198] Britnell L, Gorbachev RV, Jalil R, Belle BD, Schedin F, Mishchenko A,

Georgiou T, Katsnelson MI, Eaves L, Morozov SV, Peres NMR, Leist J, Geim

AK, Novoselov KS, Ponomarenko LA. Field-effect tunneling transistor based

on vertical graphene heterostructures. Science 2012;335:947-50.

[199] Britnell L, Gorbachev RV, Jalil R, Belle BD, Schedin F, Katsnelson MI, Eaves

L, Morozov SV, Mayorov AS, Peres NMR, Castro Neto AH, Leist J, Geim AK,

Ponomarenko LA, Novoselov KS. Electron tunneling through ultrathin boron

nitride crystalline barriers. Nano Lett 2012;12:1707-10.

[200] Dean CR, Young AF, Meric I, Lee C, Wang L, Sorgenfrei S, Watanabe K,

Taniguchi T, Kim P, Shepard KL, Hone J. Boron nitride substrates for

high-quality graphene electronics. Nat Nanotech 2010;5:722-6.

[201] Wang L, Chen Z, Dean CR, Taniguchi T, Watanabe K, Brus LE, Hone J.

Negligible environmental sensitivity of graphene in a hexagonal boron

nitride/graphene/h-BN sandwich structure. ACS Nano 2012;6:9314-9.

[202] Levendorf MP, Kim CJ, Brown L, Huang PY, Havener RW, Muller DA, Park J.

Graphene and boron nitride lateral heterostructures for atomically thin circuitry.

Nature 2012;488:627-32.

[203] Decker R, Wang Y, Brar VW, Regan W, Tsai HZ, Wu Q, Gannett W, Zettl A,

Crommie MF. Local electronic properties of graphene on a BN substrate via

scanning tunneling microscopy. Nano Lett 2011;11:2291-5.

[204] Xue J, Sanchez-Yamagishi J, Bulmash D, Jacquod P, Deshpande A, Watanabe

K, Taniguchi T, Jarillo-Herrero P, LeRoy BJ. Scanning tunnelling microscopy

Page 146: Graphene-analogous low-dimensional materials

145

and spectroscopy of ultra-flat graphene on hexagonal boron nitride. Nat Mater

2011;10:282-5.

[205] Lin T, Su C, Zhang X, Zhang W, Lee Y, Chu C, Lin H, Chang M, Chen F, Li L.

Converting graphene oxide monolayers into boron carbonitride nanosheets by

substitutional doping. Small 2012;8:1384-91.

[206] Sutter P, Cortes R, Lahiri J, Sutter E. Interface formation in monolayer

graphene-boron nitride heterostructures. Nano Lett 2012;12:4869-74.

[207] Pakdel A, Wang X, Zhi C, Bando Y, Watanabe K, Sekiguchi T, Nakayama T,

Golberg D. Facile synthesis of vertically aligned hexagonal boron nitride

nanosheets hybridized with graphitic domains. J Mater Chem 2012;22:4818-24.

[208] Kumar S, Kamaraju N, Vasu KS, Nag A, Sood AK, Rao CNR. Graphene

analogue BCN: femtosecond nonlinear optical susceptibility and hot carrier

dynamics. Chem Phys Lett 2010;499:152-7.

[209] Bhowmick S, Singh AK, Yakobson BI. Quantum dots and nanoroads of

graphene embedded in hexagonal boron nitride. J Phys Chem C

2011;115:9889-93.

[210] Kınacı A, Haskins JB, Sevik C, Cagın T. Thermal conductivity of BN-C

nanostructures. Phys Rev B 2012;86:115410.

[211] Manna AK, Pati SK. Tunable electronic and magnetic properties in BxNyCz

nanohybrids: effect of domain segregation. J Phys Chem C 2011;115:10842-50.

[ 212 ] Bernardi M, Palummo M, Grossman JC. Optoelectronic properties in

monolayers of hybridized graphene and hexagonal boron nitride. Phys Rev Lett

2012;108:226805.

[213] Jiang J, Wang J, Wang B. Minimum thermal conductance in graphene and

Page 147: Graphene-analogous low-dimensional materials

146

boron nitride superlattice. Appl Phys Lett 2011;99:043109.

[214] Zhou Y, Wang Z, Yang P, Gao F. Novel electronic and magnetic properties of

graphene nanoflakes in a boron nitride layer. J Phys Chem C 2012;116:7581-6.

[215] Sevincli H, Li W, Mingo N, Cuniberti G, Roche S. Effects of domains in

phonon conduction through hybrid boron nitride and graphene sheets. Phys Rev

B 2011;84:205444.

[ 216 ] Yang K, Chen Y, D’Agosta R, Xie Y, Zhong J, Rubio A. Enhanced

thermoelectric properties in hybrid graphene/boron nitride nanoribbons. Phys

Rev B 2012;86:045425.

[217] Nguyen VH, Mazzamuto F, Bournel A, Dollfus P. Resonant tunnelling diodes

based on graphene/h-BN heterostructure. J Phys D Appl Phys 2012;45:325104.

[218] Jungthawan S, Limpijumnong S, Kuo J. Electronic structures of graphene/boron

nitride sheet superlattices. Phys Rev B 2011;84:235424.

[219] Li J, Shenoy VB. Graphene quantum dots embedded in hexagonal boron nitride

sheets. Appl Phys Lett 2011;98:013105.

[220] Liu Y, Bhowmick S, Yakobson BI. BN white graphene with “colorful” edges:

the energies and morphology. Nano Lett 2011;11:3113-6.

[221] Zhao R, Wang J, Yang M, Liu Z, Liu Z. BN-embedded graphene with a

ubiquitous gap opening. J Phys Chem C 2012;116:21098-103.

[222] Fan X, Shen Z, Liu AQ, Kuo J. Band gap opening of graphene by doping small

boron nitride domains. Nanoscale 2012;4:2157-65.

[223] Kan E, Wu X, Li Z, Zeng XC, Yang J, Hou JG. Half-metallicity in hybrid BCN

nanoribbons. J Chem Phys 2008;129:084712.

[224] Yu Z, Hu ML, Zhang CX, He CY, Sun LZ, Zhong J. Transport properties of

Page 148: Graphene-analogous low-dimensional materials

147

hybrid zigzag graphene and boron nitride nanoribbons. J Phys Chem C

2011;115:10836-41.

[225] Liu Y, Wu X, Zhao Y, Zeng XC, Yang J. Half-Metallicity in hybrid

graphene/boron nitride nanoribbons with dihydrogenated edges. J Phys Chem C

2011;115:9442-50.

[226] Pruneda JM. Origin of half-semimetallicity induced at interfaces of C-BN

heterostructures. Phys Rev B 2010;81:161409(R).

[227] Dutta S, Manna AK, Pati SK. Intrinsic half-metallicity in modified graphene

nanoribbons. Phys Rev Lett 2009;102:096601.

[228] Lu P, Zhang Z, Guo W. Electronic structures of BC2N nanoribbons. J Phys

Chem C 2011;115:3572-7.

[229] Cao T, Feng J, Wang EG. Adsorption of hydrogen on the interface of a

graphene/boron nitride hybrid atomic membrane. Phys Rev B 2011;84:205447.

[230] Wang S, Zhang L, Xia Z, Roy A, Chang DW, Baek J, Dai L. BCN graphene as

efficient metal-free electrocatalyst for the oxygen reduction reaction. Angew

Chem Int Ed 2012;51:4209-12.

[231] Lei W, Qin S, Liu D, Portehault D, Liu Z, Chen Y. Large scale boron carbon

nitride nanosheets with enhanced lithium storage capabilities. Chem Commun

2013;49:352-4.

[232] Watanabe K, Taniguchi T, Niiyama T, Miya K, Taniguchi M. Far-ultraviolet

plane-emission handheld device based on hexagonal boron nitride. Nat

Photonics 2009;3:591-4.

[233] Zhu Y, Bando Y, Yin L, Golberg D. Field nanoemitters: ultrathin BN

nanosheets protruding from Si3N4 nanowires. Nano Lett 2006;6:2982-6.

Page 149: Graphene-analogous low-dimensional materials

148

[234] Feng PX, Sajjad M. Few-atomic-layer boron nitride sheets syntheses and

applications for semiconductor diodes. Mater Lett 2012;89:206-8.

[235] Khan U, May P, O'Neill A, Bell AP, Boussac E, Martin A, Semple J, Coleman

JN. Polymer reinforcement using liquid-exfoliated boron nitride nanosheets.

Nanoscale 2013,5:581-7.

[236] Chen Z, Zou J. Field emitters: ultrathin BN nanosheets protruded from BN

fibers. J Mater Chem 2011;21:1191-5.

[237] Terrones M, Charlier JC, Gloter A, Cruz-Silva E, Terrés E, Li YB, Vinu A,

Zanolli Z, Dominguez JM, Terrones H, Bando Y, Golberg D. Experimental and

theoretical studies suggesting the possibility of metallic boron nitride edges in

porous nanourchins. Nano Lett 2008;8:1026-32.

[238] Song W, Wang P, Cao L, Anderson A, Meziani MJ, Farr AJ, Sun Y.

Polymer/boron nitride nanocomposite materials for superior thermal transport

performance. Angew Chem Int Ed 2012;51:6498-501.

[239] Taha-Tijerina J, Narayanan TN, Gao G, Rohde M, Tsentalovich DA, Pasquali

M, Ajayan PM. Electrically insulating thermal nano-oils using 2D fillers. ACS

Nano 2012;6:1214-20.

[240] Liu Y, Xie X, Ye X. Tuning the solubility of boron nitride nanosheets in

organic solvents by using block copolymer as a “Janus” modifier. Chem

Commun 2013,49:388-90.

[241] Sainsbury T, Satti A, May P, Wang Z, McGovern I, Gun’ko YK, Coleman J.

Oxygen radical functionalization of boron nitride nanosheets. J Am Chem Soc

2012;134:18758-71.

[242] Dil H, Lobo-Checa J, Laskowski R, Blaha P, Berner S, Osterwalder J, Greber T.

Page 150: Graphene-analogous low-dimensional materials

149

Surface trapping of atoms and molecules with dipole rings. Science

2008;319:1824-6.

[243] Ding Y, Iannuzzi M, Hutter J. Investigation of boron nitride nanomesh

interacting with water. J Phys Chem C 2011;115:13685-92.

[244] Widmer R, Passerone D, Mattle T, Sachdev H, Groning O. Probing the

selectivity of a nanostructured surface by xenon adsorption. Nanoscale

2010;2:502-8.

[245] Natterer FD, Patthey F, Brune H. Ring state for single transition metal atoms on

boron nitride on Rh(111). Phys Rev Lett 2012;109:066101.

[246] Wang B, Bocquet M. Monolayer graphene and h-BN on metal substrates as

versatile templates for metallic nanoclusters. J Phys Chem Lett 2011;2:2341-5.

[247] Gao M, Lyalin A, Taketsugu T. Catalytic activity of Au and Au2 on the h-BN

surface: adsorption and activation of O2. J Phys Chem C 2012;116:9054-62.

[248] Wang L, Sun C, Xu L, Qian Y. Convenient synthesis and applications of gram

scale boron nitride nanosheets. Catal Sci Technol 2011;1:1119-23.

[249] Wang L, Shen L, Xu X, Xu L, Qian Y. Facile synthesis of uniform h-BN

nanocrystals and their application as a catalyst support towards the selective

oxidation of benzyl alcohol. RSC Adv 2012;2:10689-93.

[250] Lin Y, Bunker CE, Shiral Fernando KA, Connell JW. Aqueously dispersed

silver nanoparticle-decorated boron nitride nanosheets for reusable, thermal

oxidation-resistant surface enhanced Raman spectroscopy (SERS) devices.

ACS Appl Mater Interfaces 2012;4:1110-7.

[ 251 ] Cahangirov S, Topsakal M, Aktürk E, Şahin H, Ciraci S. Two- and

one-dimensional honeycomb structures of silicon and germanium. Phys Rev

Page 151: Graphene-analogous low-dimensional materials

150

Lett 2009;102:236804.

[252] Jose D, Datta A. An understanding of the buckling distortions in silicene. J Phys

Chem C 2012;116:24639-48.

[253] Nakano H, Mitsouka T, Harada M, Horibuchi K, Nozaki H, Takahashi N,

Nonaka T, Seno Y, Nakamura H. Soft synthesis of single-crystal silicon

monolayer sheets. Angew Chem Int Ed 2006;45:6303-6.

[254] De Padova P, Quaresima C, Olivieri B, Perfetti P, Lay GL. Strong resistance of

silicene nanoribbons towards oxidation. J Phys D Appl Phys 2011;44:312001.

[255] Davila ME, Marele A, Padova PD, Montero I, Hennies F, Pietzsch A, Shariati

MN, Gomez-Rodrıguez JM, Lay GL. Comparative structural and electronic

studies of hydrogen interaction with isolated versus ordered silicon nanoribbons

grown on Ag(110). Nanotechnology 2012;23:385703.

[256] Nakano H, Ishii M, Nakamura H. Preparation and structure of novel siloxene

nanosheets. Chem Commun 2005:2945-7.

[257] Okamoto H, Kumai Y, Sugiyama Y, Mitsuoka T, Nakanishi K, Ohta T, Nozaki

H, Yamaguchi S, Shirai S, Nakano H. Silicon nanosheets and their

self-assembled regular stacking structure. J Am Chem Soc 2010;132:2710-8.

[258] Sugiyama Y, Okamoto H, Mitsuoka T, Morikawa T, Nakanishi K, Ohta T,

Nakano H. Synthesis and optical properties of monolayer organosilicon

nanosheets. J Am Chem Soc 2010;132:5946-7.

[259] Nakano H, Nakano M, Nakanishi K, Tanaka D, Sugiyama Y, Ikuno T, Okamoto

H, Ohta T. Preparation of alkyl-modified silicon nanosheets by hydrosilylation

of layered polysilane (Si6H6). J Am Chem Soc 2012;134:5452-5.

[260] O’Hare A, Kusmartsev FV, Kugel KI. A stable “flat″ form of two-dimensional

crystals: could graphene, silicene, germanene be minigap semiconductors?

Page 152: Graphene-analogous low-dimensional materials

151

Nano Lett 2012;12:1045-52.

[261] Chen L, Liu CC, Feng B, He X, Cheng P, Ding Z, Meng S, Yao Y, Wu K.

Evidence for Dirac fermions in a honeycomb lattice based on silicon. Phys Rev

Lett 2012;109:056804.

[ 262 ] Liu CC, Feng W, Yao Y. Quantum spin hall effect in silicene and

two-dimensional germanium. Phys Rev Lett 2011;107:076802.

[263] Ni Z, Liu Q, Tang K, Zheng J, Zhou J, Qin R, Gao Z, Yu D, Lu J. Tunable

bandgap in silicene and germanene. Nano Lett 2012;12:113-8.

[264] Ezawa M. Valley-polarized metals and quantum anomalous hall effect in

silicene. Phys Rev Lett 2012;109:055502.

[265] Garcia JC, de Lima DB, Assali LVC, Justo JF. Group IV graphene- and

graphane-like nanosheets. J Phys Chem C 2011;115:13242-6.

[266] 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.

[267] Zhang C, Yan S. First-principles study of ferromagnetism in two-dimensional

silicene with hydrogenation. J Phys Chem C 2012;116:4163-6.

[268] 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.

[269] 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.

[270] Osborn TH, Farajian AA. Stability of lithiated silicene from first principles. J

Phys Chem C 2012;116:22916-20.

[271] Ding Y, Ni J. Electronic structures of silicon nanoribbons. Appl Phys Lett

Page 153: Graphene-analogous low-dimensional materials

152

2009;95:083115.

[272] Kim J, Fischetti MV, Aboud S. Structural, electronic, and transport properties

of silicane nanoribbons. Phys Rev B 2012;86:205323.

[273] An X, Zhang Y, Liu J, Li S. Spin-polarized current induced by a local exchange

field in a silicene nanoribbons. New J Phys 2012;14:083039.

[274] Xu C, Luo G, Liu Q, Zheng J, Zhang Z, Nagase S, Gao Z, Lu J. Giant

magnetoresistance in silicene nanoribbons. Nanoscale 2012;4:3111-7.

[ 275 ] Kang J, Wu F, Li J. Symmetry-dependent transport properties and

magnetoresistance in zigzag silicene nanoribbons. Appl Phys Lett

2012;100:233122.

[276] Li H, Wang L, Liu Q, Zheng J, Mei W, Gao Z, Shi J, Lu J. High performance

silicene nanoribbon field effect transistors with current saturation. Eur Phys J B

2012;85:274.

[277] Pan L, Liu HJ, Tan XJ, Lv HY, Shi J, Tang XF, Zheng G. Thermoelectric

properties of armchair and zigzag silicene nanoribbons. Phys Chem Chem Phys

2012;14:13588-93.

[278] Hu W, Wu X, Li Z, Yang J. Porous silicene as a hydrogen purification

membrane. Phys Chem Chem Phys 2013;15:5753-7.

[279] Kouvetakis J, Kaner RB, Sattler ML, Bartlett N. A novel graphite-like material

of composition BC3, and nitrogen–carbon graphites. J Chem Soc Chem

Commun 1986:1758-9.

[280] Yanagisawa H, Ishida Y, Tanaka T, Ueno A, Otani S, Oshima C. Metastable

BC3 honeycomb epitaxial sheets on the NbB2(0001) surface. Surf Sci

2006;600:4072-6.

Page 154: Graphene-analogous low-dimensional materials

153

[281] Yanagisawa H, Tanaka T, Ishida Y, Matsue M, Rokuta E, Otani S, Oshima C.

Phonon dispersion curves of a BC3 honeycomb epitaxial sheet. Phys Rev Lett

2004;93:177003.

[282] Ribeiro FJ, Cohen ML. Possible superconductivity in hole-doped BC3. Phys

Rev B 2004;69:212507.

[283] Popov IA, Boldyrev AI. Deciphering chemical bonding in a BC3 honeycomb

epitaxial sheet. J Phys Chem C 2012;116:3147-52.

[284] Ding Y, Wang Y, Ni J. Structural, electronic, and magnetic properties of defects

in the BC3 sheet from first principles. J Phys Chem C 2010;114:12416-21.

[285] Lin X, Ni J. Electronic and magnetic properties of substitutionally Fe-, Co-, and

Ni-doped BC3 honeycomb structure. J Appl Phys 2012;111:034309.

[286] Caretti I, Jimenez I. Point defects in hexagonal BN, BC3 and BCxN compounds

studied by x-ray absorption near-edge structure. J Appl Phys 2011;110:023511.

[287] Ding Y, Ni J. Tuning electronic properties of hydro-boron-carbon compounds

by hydrogen and boron contents: a first principles study. J Phys Chem C

2009;113:18468-72.

[288] Ding Y, Wang Y, Ni J, Shi L, Shi S, Li C, Tang W. First-principles study of

half-metallicity in semi-hydrogenated BC3, BC5, BC7, and B-doped graphone

sheets. Nanoscale Res Lett 2011;6:190.

[289] Ding Y, Wang Y, Ni J. Electronic structures of BC3 nanoribbons. Appl Phys

Lett 2009;94:073111.

[290] Dutta S, Wakabayashi K. Anomalous energy-gap behaviour of armchair BC3

ribbons due to enhanced π-conjugation. J Mater Chem 2012;22:20881-4.

[291] Zhang C, Alavi A. Hydrogen absorption in bulk BC3: a first-principles study. J

Page 155: Graphene-analogous low-dimensional materials

154

Chem Phys 2007;127:214704.

[292] Sha X, Cooper AC, Bailey III WH, Cheng H. Revisiting hydrogen storage in

bulk BC3. J Phys Chem C 2010;114:3260-4.

[293] Yang Z, Ni J. Li-doped BC3 sheet for high-capacity hydrogen storage. Appl

Phys Lett 2012;100:183109.

[294] Yang Z, Ni J. Hydrogen storage on calcium-decorated BC3 sheet: a first

principles study. Appl Phys Lett 2010;97:253117.

[295] Kuzubov AA, Fedorov AS, Eliseeva NS, Tomilin FN, Avramov PV, Fedorov

DG. High-capacity electrode material BC3 for lithium batteries proposed by ab

initio simulations. Phys Rev B 2012;85:195415.

[296] Lou P, Lee JY. Band structures of narrow zigzag silicon carbon nanoribbons. J

Phys Chem C 2009;113:12637-40.

[297] Gori P, Pulci O, Marsili M, Bechstedt F. Side-dependent electron escape from

graphene- and graphane-like SiC layers. Appl Phys Lett 2012;100:043110.

[298] Ding Y, Wang Y. Electronic structures of zigzag SiC nanoribbons with

asymmetric hydrogen-terminations. Appl Phys Lett 2012;101:013102.

[299] Lou P, Lee JY. Electrical control of magnetization in marrow zigzag silicon

carbon nanoribbons. J Phys Chem C 2009;113:21213-7.

[300] Xu B, Yin J, Xia YD, Wan XG, Liu ZG. Ferromagnetic and antiferromagnetic

properties of the semihydrogenated SiC sheet. Appl Phys Lett 2010;96:143111.

[301] Lu T, Liao X, Wang H, Zheng J. Tuning the indirect–direct band gap transition

of SiC, GeC and SnC monolayer in a graphene-like honeycomb structure by

strain engineering: a quasiparticle GW study. J Mater Chem 2012;22:10062-8.

[302] Bekaroglu E, Topsakal M, Cahangirov S, Ciraci S. First-principles study of

Page 156: Graphene-analogous low-dimensional materials

155

defects and adatoms in silicon carbide honeycomb structures. Phys Rev B

2010;81:075433.

[303] Lou P, Lee JY. Spin controlling in narrow zigzag silicon carbon nanoribbons by

carrier doping. J Phys Chem C 2010;114:10947-51.

[304] Costa CD, Morbec JM. Boron and nitrogen impurities in SiC nanoribbons: an

ab initio investigation. J Phys Condens Matter 2011;23:205504.

[305] Wang XQ, Wang JT. Structural stabilities and electronic properties of fully

hydrogenated SiC sheet. Phys Lett A 2011;375:2676-9.

[306] Guan J, Chen W, Zhao X, Yu G, Huang X, Sun C. Successive hydrogenation

starting from the edge(s): an effective approach to fine-tune the electronic and

magnetic behaviors of SiC nanoribbons. J Mater Chem 2012;22:24166-72.

[307] Şahin H, Cahangirov S, Topsakal M, Bekaroglu E, Akturk E, Senger RT, Ciraci

S. Monolayer honeycomb structures of group-IV elements and III-V binary

compounds: first-principles calculations. Phys Rev B 2009;80:155453.

[308] Belanzoni P, Giorgi G, Cerofolini GF, Sgamellotti A. Planar tetracoordinated

silicon in silicon carbonyl complexes: a DFT approach. J Phys Chem A

2006;110:4582-91.

[ 309 ] Zhai HJ, Kiran B, Li J, Wang LS. Hydrocarbon analogues of boron

clusters-planarity, aromaticity and antiaromaticity. Nat Mater 2003;2:827-33.

[310] Zhang LZ, Yan QB, Du SX, Su G, Gao HJ. Boron sheet adsorbed on metal

surfaces: structures and electronic properties. J Phys Chem C

2012;116:18202-6.

[311] Evans MH, Joannopoulos JD, Pantelides ST. Electronic and mechanical

properties of planar and tubular boron structures. Phys Rev B 2005;72:045434.

Page 157: Graphene-analogous low-dimensional materials

156

[312] Cabria I, Lopez MJ, Alonso JA. Density functional calculations of hydrogen

adsorption on boron nanotubes and boron sheets. Nanotechnology

2006;17:778-85.

[313] Kunstmann J, Quandt A. Broad boron sheets and boron nanotubes: an ab initio

study of structural, electronic, and mechanical properties. Phys Rev B

2006;74:035413.

[314] Szwacki NG, Sadrzadeh A, Yakobson BI. B80 fullerene: an ab initio prediction

of geometry, stability, and electronic structure. Phys Rev Lett 2007;98:166804.

[315] Tang H, Ismail-Beigi S. Novel precursors for boron nanotubes: the competition

of two-center and three-center bonding in boron sheets. Phys Rev Lett

2007;99:115501.

[ 316 ] Lau KC, Pandey R. Thermodynamic stability of novel boron sheet

configurations. J Phys Chem B 2008;112:10217-20.

[317] Tang H, Ismail-Beigi S. First-principles study of boron sheets and nanotubes.

Phys Rev B 2010;82:115412.

[318] Zope RR, Baruah T. Snub boron nanostructures: chiral fullerenes, nanotubes

and planar sheet. Chem Phys Lett 2011;501:193-6.

[319] Özdoğan C, Mukhopadhyay S, Hayami W, Güvenc ZB, Pandey R, Boustani I.

The unusually stable B100 fullerene, structural transitions in boron

nanostructures, and a comparative study of α- and γ-boron and sheets. J Phys

Chem C 2010;114:4362-75.

[320] Penev ES, Bhowmick S, Sadrzadeh A, Yakobson BI. Polymorphism of

two-dimensional boron. Nano Lett 2012;12:2441-5.

[321] Yu X, Li L, Xu XW, Tang CC. Prediction of two-dimensional boron sheets by

Page 158: Graphene-analogous low-dimensional materials

157

particle swarm optimization algorithm. J Phys Chem C 2012;116:20075-9.

[322] Wu X, Dai J, Zhao Y, Zhuo Z, Yang J, Zeng XC. Two-dimensional boron

monolayer sheets. ACS Nano 2012;6:7443-53.

[323] Saxena S, Tyson TA. Insights on the atomic and electronic structure of boron

nanoribbons. Phys Rev Lett 2010;104:245502.

[324] Er S, de Wijs GA, Brocks G. DFT study of planar boron sheets: a new template

for hydrogen storage. J Phys Chem C 2009;113:18962-7.

[325] Wu X, Pei Y, Zeng XC. B2C graphene, nanotubes, and nanoribbons. Nano Lett

2009;9:1577-82.

[326] Dai J, Li Z, Yang J, Hou J. A first-principles prediction of two-dimensional

superconductivity in pristine B2C single layers. Nanoscale 2012;4:3032-5.

[327] An H, Liu C, Zeng Z, Fan C, Ju X. Li-doped B2C graphene as potential

hydrogen storage medium. Appl Phys Lett 2011;98:173101.

[328] Radisavljevic B, Radenovic A, Brivio J, Giacometti V, Kis A. Single-layer

MoS2 transistors. Nat Nanotech 2011;6:147-50.

[329] Brivio J, Alexander DTL, Kis A. Ripples and layers in ultrathin MoS2

membranes. Nano Lett 2011;11:5148-53.

[330] Böker Th, Severin R, Muller A, Janowitz C, Manzke R, Voss D, Kruger P,

Mazur A, Pollmann J. Band structure of MoS2, MoSe2, and α-MoTe2:

angle-resolved photoelectron spectroscopy and ab initio calculations. Phys Rev

B 2001;64:235305.

[331] Klein A, Tiefenbacher S, Eyert V, Pettenkofer C, Jaegermann W. Electronic

band structure of single-crystal and single-layer WS2: influence of interlayer

van der Waals interactions. Phys Rev B 2001;64:205416.

Page 159: Graphene-analogous low-dimensional materials

158

[332] Ataca C, Topsakal M, Aktürk E, Ciraci S. A comparative study of lattice

dynamics of three- and two-dimensional MoS2. J Phys Chem C

2011;115:16354-61.

[333] Lebègue S, Eriksson O. Electronic structure of two-dimensional crystals from

ab initio theory. Phys Rev B 2009;79:115409.

[ 334 ] Johari P, Shenoy VB. Tunable dielectric properties of transition metal

dichalcogenides. ACS Nano 2011;5:5903-8.

[335] Andersen A, Kathmann SM, Lilga MA, Albrecht KO, Hallen RT, Mei D.

First-principles characterization of potassium intercalation in hexagonal

2H-MoS2. J Phys Chem C 2012;116:1826-32.

[336] Mak KF, Lee C, Hone J, Shan J, Heinz TF. Atomically thin MoS2: a new

direct-gap semiconductor. Phys Rev Lett 2010;105:136805.

[337] Kuc A, Zibouche N, Heine T. Influence of quantum confinement on the

electronic structure of the transition metal sulfide TS2. Phys Rev B

2011;83:245213.

[338] Splendiani A, Sun L, Zhang Y, Li T, Kim J, Chim CY, Galli G, Wang F.

Emerging photoluminescence in monolayer MoS2. Nano Lett 2010;10:1271-5.

[ 339 ] Eda G, Yamaguchi H, Voiry D, Fujita T, Chen M, Chhowalla M.

Photoluminescence from chemically exfoliated MoS2. Nano Lett

2011;11:5111-6.

[340] Ma Y, Dai Y, Guo M, Niu C, Lu J, Huang B. Electronic and magnetic

properties of perfect, vacancy-doped, and nonmetal adsorbed MoSe2, MoTe2

and WS2 monolayers. Phys Chem Chem Phys 2011;13:15546-53.

[341] Voß D, Krüger P, Mazur A, Pollmann J. Atomic and electronic structure of

Page 160: Graphene-analogous low-dimensional materials

159

WSe2 from ab initio theory: bulk crystal and thin film systems. Phys Rev B

1999;60:14311-7.

[342] Tongay S, Zhou J, Ataca C, Lo K, Matthews TS, Li J, Grossman JC, Wu J.

Thermally driven crossover from indirect toward direct bandgap in 2D

semiconductors: MoSe2 versus MoS2. Nano Lett 2012;12:5576-80.

[343] Xiao D, Liu G, Feng W, Xu X, Yao W. Coupled spin and valley physics in

monolayers of MoS2 and other group-VI dichalcogenides. Phys Rev Lett

2012;108:196802.

[344] Cao T, Wang G, Han W, Ye H, Zhu C, Shi J, Niu Q, Tan P, Wang E, Liu B,

Feng J. Valley-selective circular dichroism of monolayer molybdenum

disulphide. Nat Commun 2012;3:887.

[345] Mak KF, He K, Shan J, Heinz TF. Control of valley polarization in monolayer

MoS2 by optical helicity. Nat Nanotech 2012;7:494-8.

[346] Zeng H, Dai J, Yao W, Xiao D, Cui X. Valley polarization in MoS2 monolayers

by optical pumping. Nat Nanotech 2012;7:490-3.

[347] Castellanos-Gomez A, Poot M, Steele GA, van der Zant HSJ, Agraït N,

Rubio-Bollinger G. Elastic properties of freely suspended MoS2 nanosheets.

Adv Mater 2012;24:772-5.

[348] Bertolazzi S, Brivio J, Kis A. Stretching and breaking of ultrathin MoS2. ACS

Nano 2011;5:9703-9.

[349] Scalise E, Houssa M, Pourtois G, Afanas’ev V, Stesmans A. Strain-induced

semiconductor to metal transition in the two-dimensional honeycomb structure

of MoS2. Nano Res 2012;5:43-8.

Page 161: Graphene-analogous low-dimensional materials

160

[350] Johari P, Shenoy VB. Tuning the electronic properties of semi-conducting

transition metal dichalcogenides by applying mechanical strains. ACS Nano

2012;6:5449-56.

[351] Liu Q, Li L, Li Y, Gao Z, Chen Z, Lu J. Tuning electronic structure of bilayer

MoS2 by vertical electric field: a first-principles investigation. J Phys Chem C

2012;116:21556-62.

[352] He J, Wu K, Sa R, Li Q, Wei Y. Magnetic properties of nonmetal atoms

absorbed MoS2 monolayers. Appl Phys Lett 2010;96:082504.

[353] Ataca C, Ciraci S. Functionalization of single-layer MoS2 honeycomb structures.

J Phys Chem C 2011;115:13303-11.

[354] Liu Z, Suenaga K, Wang Z, Shi Z, Okunishi E, Iijima S. Identification of active

atomic defects in a monolayered tungsten disulphide nanoribbons. Nat

Commun 2011;2:213.

[355] Hansen LP, Ramasse QM, Kisielowski C, Brorson M, Johnson E, Topsøe H,

Helveg S. Atomic-scale edge structures on industrial-style MoS2 nanocatalysts.

Angew Chem Int Ed 2011;50:10153-6.

[356] Komsa H, Kotakoski J, Kurasch S, Lehtinen O, Kaiser U, Krasheninnikov AV.

Two-dimensional transition metal dichalcogenides under electron irradiation:

defect production and doping. Phys Rev Lett 2012;109:035503.

[357] Ataca C, Ciraci S. Dissociation of H2O at the vacancies of single-layer MoS2.

Phys Rev B 2012;85:195410.

[358] Zou X, Liu Y, Yakobson BI. Predicting dislocations and grain boundaries in

two-dimensional metal-disulfides from the first principles. Nano Lett

2013;13:253-8.

Page 162: Graphene-analogous low-dimensional materials

161

[359] Ivanovskaya VV, Zobelli A, Gloter A, Brun N, Serin V, Colliex C. Ab initio

study of bilateral doping within the MoS2-NbS2 system. Phys Rev B

2008;78:134104.

[ 360 ] Komsa HP, Krasheninnikov AV. Two-dimensional transition metal

dichalcogenide alloys: stability and electronic properties. J Phys Chem Lett

2012;3:3652-6.

[361] Zhang J, Soon JM, Loh KP, Yin J, Ding J, Sullivian MB, Wu P. Magnetic

molybdenum disulfide nanosheet films. Nano Lett 2007;7:2370-6.

[362] Tongay S, Varnoosfaderani SS, Appleton BR, Wu J, Hebard AF. Magnetic

properties of MoS2: existence of ferromagnetism. Appl Phys Lett

2012;101:123105.

[363] Ataca C, Şahin H, Aktürk E, Ciraci S. Mechanical and electronic properties of

MoS2 nanoribbons and their defects. J Phys Chem C 2011;115:3934-41.

[364] Kou L, Tang C, Zhang Y, Heine T, Chen C, Frauenheim T. Tuning magnetism

and electronic phase transitions by strain and electric field in zigzag MoS2

nanoribbons. J Phys Chem Lett 2012;3:2934-41.

[365] Dolui K, Pemmaraju CD, Sanvito S. Electric field effects on armchair MoS2

nanoribbons. ACS Nano 2012;6:4823-34.

[366] Wang R, Ruzicka BA, Kumar N, Bellus MZ, Chiu HY, Zhao H. Ultrafast and

spatially resolved studies of charge carriers in atomically thin molybdenum

disulfide. Phys Rev B 2012;86:045406.

[ 367 ] Fang H, Chuang S, Chang TC, Takei K, Takahashi T, Javey A.

High-performance single layered WSe2 p-FETs with chemically doped contacts.

Nano Lett 2012;12:3788-92.

Page 163: Graphene-analogous low-dimensional materials

162

[368] Liu L, Kumar SB, Ouyang Y, Guo J. Performance limits of monolayer

transition metal dichalcogenide transistors. IEEE Trans Electron Devices

2011;58:3042-7.

[369] Ghatak S, Pal AN, Ghosh A. Nature of electronic states in atomically thin MoS2

field-effect transistors. ACS Nano 2011;5:7707-12.

[370] Li H, Yin Z, He Q, Li H, Huang X, Lu G, Fam DWH, Tok ALY, Zhang Q,

Zhang H. Fabrication of single- and multilayer MoS2 film-based field-effect

transistors for sensing NO at room temperature. Small 2012;8:63-7.

[371] Zhang Y, Ye J, Matsuhashi Y, Iwasa Y. Ambipolar MoS2 thin flake transistors.

Nano Lett 2012;12:1136-40.

[372] Braga D, Lezama IG, Berger H, Morpurgo AF. Quantitative determination of

the band gap of WS2 with ambipolar ionic liquid-gated transistors. Nano Lett

2012;12:5218-23.

[373] Kim S, Konar A, Hwang WS, Lee JH, Lee J, Yang J, Jung C, Kim H, Yoo JB,

Choi JY, Jin YW, Lee SY, Jena D, Choi W, Kim K. High-mobility and

low-power thin-film transistors based on multilayer MoS2 crystals. Nat

Commun 2012;3:1011.

[374] Late DJ, Liu B, Matte HSSR, Dravid VP, Rao CNR. Hysteresis in single-layer

MoS2 field effect transistors. ACS Nano 2012;6:5635-41.

[375] Hwang WS, Remskar M, Yan R, Protasenko V, Tahy K, Chae SD, Zhao P,

Konar A, Xing H, Seabaugh A, Jena D. Transistors with chemically synthesized

layered semiconductor WS2 exhibiting 105 room temperature modulation and

ambipolar behavior. Appl Phys Lett 2012;101:013107.

[376] Pu J, Yomogida Y, Liu KK, Li LJ, Iwasa Y, Takenobu T. Highly flexible MoS2

Page 164: Graphene-analogous low-dimensional materials

163

thin-film transistors with ion gel dielectrics. Nano Lett 2012;12:4013-7.

[377] Yin Z, Li H, Li H, Jiang L, Shi Y, Sun Y, Lu G, Zhang Q, Chen X, Zhang H.

Single-layer MoS2 phototransistors. ACS Nano 2012;6:74-80.

[378] Lee HS, Min SW, Chang YG, Park MK, Nam T, Kim H, Kim JH, Ryu S, Im S.

MoS2 nanosheet phototransistors with thickness-modulated optical energy gap.

Nano Lett 2012;12:3695-700.

[379] Radisavljevic B, Whitwick MB, Kis A. Integrated circuits and logic operations

based on single-layer MoS2. ACS Nano 2011;5:9934-8.

[380] Wang H, Yu L, Lee YH, Shi Y, Hsu A, Chin ML, Li LJ, Dubey M, Kong J,

Palacios T. Integrated circuits based on bilayer MoS2 transistors. Nano Lett

2012;12:4674-80.

[381] He Q, Zeng Z, Yin Z, Li H, Wu S, Huang X, Zhang H. Fabrication of flexible

MoS2 thin-film transistor arrays for practical gas-sensing applications. Small

2012;8:2994-9.

[382] Wu S, Zeng Z, He Q, Wang Z, Wang SJ, Du Y, Yin Z, Sun X, Chen W, Zhang

H. Electrochemically reduced single-layer MoS2 nanosheets: characterization,

properties, and sensing applications. Small 2012;8:2264-70.

[383] Xiao J, Choi D, Cosimbescu L, Koech P, Liu J, Lemmon JP. Exfoliated MoS2

nanocomposite as an anode material for lithium ion batteries. Chem Mater

2010;22:4522-4.

[384] Ding S, Chen JS, Lou XW. Glucose-assisted growth of MoS2 nanosheets on

CNT backbone for improved lithium storage properties. Chem Eur J

2011;17:13142-5.

[385] Chang K, Chen W. Single-layer MoS2/graphene dispersed in amorphous carbon:

Page 165: Graphene-analogous low-dimensional materials

164

towards high electrochemical performances in rechargeable lithium ion

batteries. J Mater Chem 2011;21:17175-84.

[386] Li YF, Wu DH, Zhou Z, Cabrera CR, Chen ZF. Enhanced Li adsorption and

diffusion on MoS2 zigzag nanoribbons by edge effects-a computational study. J

Phys Chem Lett 2012;3:2221-7.

[387] Shanmugam M, Bansal T, Durcan CA, Yu B. Molybdenum disulphide/titanium

dioxide nanocomposite-poly 3-hexylthiophene bulk heterojunction solar cell.

Appl Phys Lett 2012;100:153901.

[388] Shanmugam M, Durcan CA, Yu B. Layered semiconductor molybdenum

disulfide nanomembrane based Schottky-barrier solar cells. Nanoscale

2012;4:7399-405.

[389] Ma Y, Dai Y, Wei W, Niu C, Yu L, Huang B. First-principles study of the

graphene@MoSe2 heterobilayers. J Phys Chem C 2011;115:20237-41.

[390] Liu B, Han Y, Gao C, Ma Y, Peng G, Wu B, Liu C, Wang Y, Hu T, Cui X, Ren

W, Li Y, Su N, Liu H, Zou G. Pressure induced semiconductor-semimetal

transition in WSe2. J Phys Chem C 2010;114:14251-4.

[391] Ma Y, Dai Y, Guo M, Niu C, Zhu Y, Huang B. Evidence of the existence of

magnetism in pristine VX2 monolayers (X = S, Se) and their strain-induced

tunable magnetic properties. ACS Nano 2012;6:1695-701.

[392] Zhou Y, Wang Z, Yang P, Zu X, Yang L, Sun X, Gao F. Tensile strain switched

ferromagnetism in layered NbS2 and NbSe2. ACS Nano 2012;6:9727-36.

[393] Sun Y, Cheng H, Gao S, Sun Z, Liu Q, Liu Q, Lei F, Yao T, He J, Wei S, Xie Y.

Freestanding tin disulfide single-layers realizing efficient visible-light water

splitting. Angew Chem Int Ed 2012;51:8727-31.

Page 166: Graphene-analogous low-dimensional materials

165

[394] Zhang H, Liu CX, Qi XL, Dai X, Fang Z, Zhang SC. Topological insulators in

Bi2Se3, Bi2Te3, and Sb2Te3 with a single Dirac cone on the surface. Nat Phys

2009;5:438-42.

[ 395 ] Ghaemi P, Mong RSK, Moore JE. In-plane transport and enhanced

thermoelectric performance in thin films of the topological insulators Bi2Te3

and Bi2Se3. Phys Rev Lett 2010;105:166603.

[396] Steinberg H, Gardner DR, Lee YS, Herrero PJ. Surface state transport and

ambipolar electric field effect in Bi2Se3 nanodevices. Nano Lett

2010;10:5032-6.

[ 397 ] Ma R, Sasaki T. Nanosheets of oxides and hydroxides: ultimate 2D

charge-bearing functional crystallites. Adv Mater 2010;22:5082-104.

[398] Ida S, Shiga D, Koinuma M, Matsumoto Y. Synthesis of hexagonal nickel

hydroxide nanosheets by exfoliation of layered nickel hydroxide intercalated

with dodecyl sulfate ions. J Am Chem Soc 2008;130:14038-9.

[399] Sakai N, Ebina Y, Takada K, Sasaki T. Electronic band structure of titania

semiconductor nanosheets revealed by electrochemical and

photoelectrochemical studies. J Am Chem Soc 2004;126:5851-8.

[ 400 ] Osada M, Ebina Y, Fukuda K, Ono K, Takada K, Yamaura K,

Takayama-Muromachi E, Sasaki T, Ferromagnetism in two-dimensional

Ti0.8Co0.2O2 nanosheets. Phys Rev B 2006;73:153301.

[401] Omomo Y, Sasaki T, Wang L, Watanabe M. Redoxable nanosheet crystallites

of MnO2 derived via delamination of a layered manganese oxide. J Am Chem

Soc 2003;125:3568-75.

Page 167: Graphene-analogous low-dimensional materials

166

[402] Fukuda K, Saida T, Sato J, Yonezawa M, Takasu Y, Sugimoto W. Synthesis of

nanosheet crystallites of ruthenate with an α-NaFeO2-related structure and its

electrochemical supercapacitor property. Inorg Chem 2010;49:4391-3.

[403] Fukuda K, Akatsuka K, Ebina Y, Ma R, Takada K, Nakai I, Sasaki T.

Exfoliated nanosheet crystallite of cesium tungstate with 2D pyrochlore

structure: synthesis, characterization, and photochromic properties. ACS Nano

2008;2: 1689-95.

[404] Oh JM, Biswick TT, Choy JH. Layered nanomaterials for green materials. J

Mater Chem 2009;19:2553–63.

[405] Sels B, Vos DD, Buntinx M, Pierard F, Kirsch-De Mesmaeker A, Jacobs P.

Layered double hydroxides exchanged with tungstate as biomimetic catalysts

for mild oxidative bromination. Nature 1999;400:855-7.

[406] Liu Z, Ma R, Osada M, Iyi N, Ebina Y, Takada K, Sasaki T. Synthesis, anion

exchange, and delamination of Co-Al layered double hydroxide: assembly of

the exfoliated nanosheet/polyanion composite films and magneto-optical

studies. J Am Chem Soc 2006;128:4872-80.

[407] Hu L, Ma R, Ozawa TC, Sasaki T. Exfoliation of layered europium hydroxide

into unilamellar nanosheets. Chem Asian J 2010;5:248-51.

[408] Londero E, Schröder E. Role of van der Waals bonding in the layered oxide

V2O5: first-principles density-functional calculations. Phys Rev B

2010;82:054116.

[409] Tang Q, Li F, Zhou Z, Chen Z. Versatile electronic and magnetic properties of

corrugated V2O5 two-dimensional crystal and its derived one-dimensional

nanoribbons: a computational exploration. J Phys Chem C 2011;115:11983-90.

Page 168: Graphene-analogous low-dimensional materials

167

[410] Du AJ, Smith SC, Yao XD, Sun CH, Li L, Lu GQ. The role of V2O5 on the

dehydrogenation and hydrogenation in magnesium hydride: an ab initio study.

Appl Phys Lett 2008;92:163106.

[411] Claeyssens F, Freeman CL, Allan NL, Sun Y, Ashfold MNR, Harding JH.

Growth of ZnO thin films—experiment and theory. J Mater Chem

2005;15:139-48.

[412] Freeman CL, Claeyssens F, Allan NL. Graphitic nanofilms as precursors to

wurtzite films: theory. Phys Rev Lett 2006;96:066102.

[413] Tusche C, Meyerheim HL, Kirschner J. Observation of depolarized ZnO(0001)

monolayers: formation of unreconstructed planar sheets. Phys Rev Lett

2007;99:026102.

[414] Topsakal M, Cahangirov S, Bekaroglu E, Ciraci S. First-principles study of zinc

oxide honeycomb structures. Phys Rev B 2009;80:235119.

[415] Li H, Dai J, Li J, Zhang S, Zhou J, Zhang L, Chu W, Chen D, Zhao H, Yang J,

Wu Z. Electronic structures and magnetic properties of GaN sheets and

nanoribbons. J Phys Chem C 2010;114:11390-4.

[416] Tang Q, Cui Y, Li Y, Zhou Z, Chen Z. How do surface and edge effects alter

the electronic properties of GaN nanoribbons? J Phys Chem C

2011;115:1724-31.

[417] Du AJ, Zhu ZH, Chen Y, Lu GQ, Smith SC. First principle studies of zigzag

AlN nanoribbon. Chem Phys Lett 2009;469:183-5.

[418] Wu W, Lu P, Zhang Z, Guo W. Electronic and magnetic properties and

structural stability of BeO sheet and nanoribbons. ACS Appl Mater Interfaces

2011;3:4787-95.

Page 169: Graphene-analogous low-dimensional materials

168

[419] Zhang YG, He HY, Pan BC. Structural features and electronic properties of

MgO nanosheets and nanobelts. J Phys Chem C 2012;116:23130-5.

[420] Tu ZC. First-principles study on physical properties of a single ZnO monolayer

with graphene-like structure. J Comput Theory Nanosci 2010;7:1182-6.

[421] Schmidt TM, Miwa RH, Fazzio A. Ferromagnetic coupling in a Co-doped

graphenelike ZnO sheet. Phys Rev B 2010;81:195413.

[422] Guo H, Zhao Y, Lu N, Kan E, Zeng XC, Wu X, Yang J. Tunable magnetism in

a monmetal-substituted ZnO monolayer: a first-principles study. J Phys Chem

C 2012;116:11336-42.

[423] Tang Q, Li Y, Zhou Z, Chen Y, Chen Z. Tuning electronic and magnetic

properties of wurtzite ZnO nanosheets by surface hydrogenation. ACS Appl

Mater Interfaces 2010;2:2442-7.

[424] Chen Q, Wang J, Zhu L, Wang S, Ding F. Fluorination induced half metallicity

in two-dimensional few zinc oxide layers. J Chem Phys 2010;132:204703.

[425] Kou L, Li C, Zhang Z, Guo W. Electric-field- and

hydrogen-passivation-induced band modulations in armchair ZnO nanoribbons.

J Phys Chem C 2010;114:1326-30.

[426] Kou L, Li C, Zhang Z, Guo W. Tuning magnetism in zigzag ZnO nanoribbons

by transverse electric fields. ACS Nano 2010;4:2124-8.

[427] Botello-Méndez AR, López-Urías F, Terrones M, Terrones H. Enhanced

ferromagnetism in ZnO nanoribbons and clusters passivated with sulfur. Nano

Res 2008;1:420-6.

[428] Chen Q, Zhu L, Wang J. Edge-passivation induced half-metallicity of zigzag

zinc oxide nanoribbons. Appl Phys Lett 2009;95:133116.

Page 170: Graphene-analogous low-dimensional materials

169

[429] Si H, Pan BC. Strain-induced semiconducting-metallic transition for ZnO

zigzag nanoribbons. J Appl Phys 2010;107:094313.

[430] Botello-Méndez AR, López-Urías F, Terrones M, Terrones H. Magnetic

behavior in zinc oxide zigzag nanoribbons. Nano Lett 2008;8:1562-5.

[431] Barsoum MW. The MN+1AXN phases: a new class of solids. Prog Solid State

Chem 2000;28:201-81.

[432] Naguib M, Kurtoglu M, Presser V, Lu J, Niu J, Heon M, Hultman L, Gogotsi Y,

Barsoum MW. Two-dimensional nanocrystals produced by exfoliation of

Ti3AlC2. Adv Mater 2011;23:4248-53.

[433] Naguib M, Mashtalir O, Carle J, Presser V, Lu J, Hultman L, Gogotsi Y,

Barsoum MW. Two-dimensional transition metal carbides. ACS Nano

2012;6:1322-31.

[434] Tang Q, Zhou Z, Shen P. Are MXenes promising anode materials for Li ion

batteries? Computational studies on electronic properties and Li storage

capability of Ti3C2 and Ti3C2X2 (X = F, OH) monolayer. J Am Chem Soc

2012;134:16909-16.

[435] Khazaei M, Arai M, Sasaki T, Chung C, Venkataramanan NS, Estili M, Sakka

Y, Kawazoe Y. Novel electronic and magnetic properties of two-dimensional

transition metal carbides and nitrides. Adv Funct Mater DOI:

10.1002/adfm.201202502.

[436] Naguib M, Come J, Dyatkin B, Presser V, Taberna PL, Simon P, Barsoum MW,

Gogotsi Y. MXene: a promising transition metal carbide anode for lithium-ion

batteries. Electrochem Commun 2012;16:61-4.

[437] Come J, Naguid M, Rozier P, Barsoum MW, Gogotsi Y, Taberna PL, Morcrette

Page 171: Graphene-analogous low-dimensional materials

170

M, Simon P. A non-aqueous asymmetric cell with a Ti2C-based

two-dimensional negative electrode. J Electrochem Soc 2012;159:A1368-73.

[438] Tang Q, Zhou Z, Chen Z. Single-layer [Cu2Br(IN)2]n coordination polymer (CP):

electronic and magnetic properties, and implication for molecular sensors. J

Phys Chem C 2012;116:4119-25.

[439] Li Y, Xiao J, Shubina TE, Chen M, Shi Z, Schmid M, Steinrück HP, Gottfried

JM, Lin N. Coordination and metalation bifunctionality of Cu with

5,10,15,20-tetra(4-pyridyl)porphyrin: toward a mixed-valence two-dimensional

coordination network. J Am Chem Soc 2012;134:6401-8.

[440] Zhou J, Sun Q. Magnetism of phthalocyanine-based organometallic single

porous sheet. J Am Chem Soc 2011;133:15113-9.

[441] Bauer T, Zheng Z, Renn A, Enning R, Stemmer A, Sakamoto J, Schlüter AD.

Synthesis of free-standing, monolayered organometallic sheets at the air/water

interface. Angew Chem Int Ed 2011;50:7879-84.

[442] Shi Z, Liu J, Lin T, Xia F, Liu PN, Lin N. Thermodynamics and selectivity of

two-dimensional metallo-supramolecular self-assembly resolved at molecular

scale. J Am Chem Soc 2011;133:6150-3.

[443] Kan E, Wu X, Lee C, Shim JH, Lu R, Xiao C, Deng K. Two-dimensional

organometallic porous sheets with possible high-temperature ferromagnetism.

Nanoscale 2012;4:5304-7.

[444] Spitler EL, Dichtel WR. Lewis acid-catalysed formation of two-dimensional

phthalocyanine covalent organic frameworks. Nature Mater 2010;2:672-7.

[445] Uribe-Romo FJ, Hunt JR, Furukawa H, Klöck C, O’Keeffe M, Yaghi OM. A

crystalline imine-linked 3-D porous covalent organic framework. J Am Chem

Page 172: Graphene-analogous low-dimensional materials

171

Soc 2009;131:4570-1.

[446] Kuhn P, Antonietti M, Thomas A. Porous, covalent triazine-based frameworks

prepared by ionothermal synthesis. Angew Chem Int Ed 2008;47:3450-3.

[447] Hunt JR, Doonan CJ, LeVangie JD, Côté AP, Yaghi OM. Reticular synthesis of

covalent organic borosilicate frameworks. J Am Chem Soc 2008;130:11872-3.

[448] Uribe-Romo FJ, Doonan CJ, Furukawa H, Oisaki K, Yaghi OM. Crystalline

covalent organic frameworks with hydrazone linkages. J Am Chem Soc

2011;133:11478-81.

[449] Bieri M, Nguyen MT, Gröning O, Cai J, Treier M, Aït-Mansour K, Ruffieux P,

Pignedoli CA, Passerone D, Kastler M, Müllen K, Fasel R. Two-dimensional

polymer formation on surfaces: insight into the roles of precursor mobility and

reactivity. J Am Chem Soc 2010;132:16669-76.

[450] Zwaneveld NAA, Pawlak R, Abel M, Catalin D, Gigmes D, Bertin D, Porte L.

Organized formation of 2D extended covalent organic frameworks at surfaces. J

Am Chem Soc 2008;130:6678-9.

[451] Clair S, Ourdjini O, Abel M, Porte L. Tip- or electron beam-induced surface

polymerization. Chem Commun 2011;47:8028-30.

[452] Colson JW, Woll AR, Mukherjee A, Levendorf MP, Spitler EL, Shields VB,

Spencer MG, Park J, Dichtel WR. Oriented 2D covalent organic framework

thin films on single-layer graphene. Science 2011;332:228-31.

[453] Zhou Y, Wang Z, Yang P, Zu X, Gao F. Electronic and optical properties of

two-dimensional covalent organic frameworks. J Mater Chem

2012;22:16964-70.

[454] Bieri M, Blankenburg S, Kivala M, Pignedoli CA, Ruffieux P, Müllen K, Fasel

Page 173: Graphene-analogous low-dimensional materials

172

R. Surface-supported 2D heterotriangulene polymers. Chem Commun

2011;47:10239-41.

[455] Kan E, Hu W, Xiao C, Lu R, Deng K, Yang J, Su H. Half-metallicity in organic

single porous sheets. J Am Chem Soc 2012;134:5718-21.


Recommended