0% found this document useful (0 votes)
10 views29 pages

2D Layered Transition Metal Dichalcogenides (Mos) : Synthesis, Applications and Theoretical Aspects

Uploaded by

Neha
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
10 views29 pages

2D Layered Transition Metal Dichalcogenides (Mos) : Synthesis, Applications and Theoretical Aspects

Uploaded by

Neha
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 29

Applied Materials Today 13 (2018) 242–270

Contents lists available at ScienceDirect

Applied Materials Today


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

Review

2D layered transition metal dichalcogenides (MoS2 ): Synthesis, applications and


theoretical aspects
Arun Kumar Singh a,1 , P. Kumar b,1 , D.J. Late b , Ashok Kumar c , S. Patel d , Jai Singh d,∗
a
Department of Physics, Motilal Nehru National Institute of Technology, Allahabad 211004, India
b
Physical & Materials Chemistry Division, National Chemical Laboratory, Pashan Road, Pune 411008, India
c
Department of Physical Sciences, School of Basic and Applied Sciences, Central University of Punjab, Bathinda 151001, India
d
Department of Physics, Dr. H.S.G. University, Sagar, 470003, India

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

Article history: Recently, graphene and other two-dimensional (2D) transition metal dichalcogenides (TMDCs) have been
Received 10 July 2018 widely explored due to their unique optical, mechanical, electrical and sensing properties for versatile
Received in revised form 7 September 2018 electronic and optoelectronic applications. The atomically thin layers of TMDC materials have shown
Accepted 10 September 2018
potential to replace state-of-the-art silicon-based technology. Graphene has already revealed an excess
of new physics and multifaceted applications in several areas. Similarly, mono-layers of TMDCs such
Keywords:
as molybdenum disulfide (MoS2 ) have also shown excellent electrical and optical properties possessing
Two-dimensional materials
a direct band-gap of ∼1.8 eV combined with high mechanical flexibility. In contrast to semi-metallic
Transition metal dichalcogenides
MoS2
graphene, the semiconducting behavior of MoS2 allows it to overcome the deficiencies of zero-band-gap
Transistors graphene. This review summarizes the synthesis of 2D MoS2 by several techniques, i.e., mechanical and
Sensors chemical exfoliation, RF-sputtering, atomic layer deposition (ALD) and chemical vapor deposition (CVD),
Solar cell etc. Furthermore, extensive studies based on potential applications of MoS2 such as the sensor, solar
Field emission cells, field emission and as an efficient catalyst for hydrogen generation has been included. Theoretical
Topological properties aspects combined with the experimental observations to provide more insights on the dielectric, optical
and topological behavior of MoS2 was highlighted.
© 2018 Published by Elsevier Ltd.

Contents

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 243
1.1. Benefits of MoS2 over Silicon . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 243
2. Properties of transition metal dichalcogenides . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 244
2.1. Crystal structures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 244
2.2. Electronic band structures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 244
2.3. Electrical and optical properties . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 245
2.4. Vibrational properties . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 247
2.5. Mechanical properties . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 247
3. Theoretical aspects . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 249
3.1. First principles perspective of MoS2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 249
3.2. Background . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 250
3.3. Electronic structure of MoS2 as a function of layers thickness . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 251
3.4. Dielectric response of MoS2 : electron energy loss spectroscopy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 252
3.5. A topological aspect of MoS2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 253
3.6. Topological Z2 invariant . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 254
3.7. 1T’-MoS2 as topological insulators . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 254

∗ Corresponding author.
E-mail address: [email protected] (J. Singh).
1
These authors contributed equally to this work.

https://doi.org/10.1016/j.apmt.2018.09.003
2352-9407/© 2018 Published by Elsevier Ltd.
A.K. Singh et al. / Applied Materials Today 13 (2018) 242–270 243

4. Growth methods . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 255


4.1. Mechanical exfoliation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 255
4.2. Chemical vapor deposition . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 255
4.2.1. Vapor phase growth . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 256
4.3. Liquid-phase preparations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 257
4.4. MoS2 film fabricated by RF sputtering . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 258
4.5. Atomic layer deposition (ALD) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 258
5. Applications . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 260
5.1. TMDCs transistors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 260
5.2. Field emission . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 261
5.3. Photovoltaics and photodetection . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 261
5.4. Gas sensing . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 262
5.5. Hydrogen production . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 262
6. Hydrogen evolution mechanism . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 264
7. Composites with conductive hosts . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 265
8. Defective MoS2 nanosheets and amorphous MoS2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 267
9. Role of the phase, edge, and vacancies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 267
10. The photoelectrocatalytic (PEC) HER . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 267
11. Prospects and direction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 267
12. Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 268
Acknowledgments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 268
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 268

1. Introduction Usually, bulk counterparts of the multilayered materials are


formed by the weak interaction between the layers, which serves
The discovery of graphene in 2004 by Geim & Novoselov as to stabilize the structure. TMDCs (e.g., MoS2 , WS2 , WSe2 , etc.) con-
the first two-dimensional (2D) material with excellent electronic, sist of a sandwich structure in which a layer of transition metal
optical and mechanical properties has given birth to research (e.g., Mo, W, Nb) atoms are hexagonally packed between two lay-
on several other 2D materials [1]. Graphene has been exten- ers of chalcogen (e.g., S, Se, Te) atom. The intercalation or doping
sively investigated due to its unique features such as high field into the interlamellar space of these materials enables scientists to
effect mobility ∼104 –105 cm2 V−1 s−1 , superior flexibility, bet- tailor their physical properties. As the number of layers is closely
ter transparency, improved chemical stability, and high surface related to the physical properties, thus allowing to tune electri-
area [1,2]. Later-on, emerged as a promising candidate for the cal properties by adjusting the number of layers. With the help of
post-silicon age with numerous technological applications in nano- the accumulated synthetic knowledge, there has been particular
electronics, optoelectronics, spintronics, energy harvesting and interest in monolayered materials.
sensors [2c–e,3]. However, pristine graphene has no band gap,
so it is unlikely to be utilized for the fabrication of logical cir- 1.1. Benefits of MoS2 over Silicon
cuits. In this context, 2D materials, especially TMDCs, have gained
worldwide attention in recent years for possible application in TMDCs have certain advantages over silicon (Si) when it comes
field-effect transistors (FETs), integrated circuits, photodetectors, to creating a microchip for miniaturization. In a 0.65 nm thin sheet
memory devices, chemical and biosensors, lithium-ion batter- of MoS2 , the electrons can move around as quickly as in a 2 nm thick
ies, hydrogen evolution catalysis, and supercapacitors [3d,e,4,5]. sheet of Si. It is not currently possible to fabricate a layer of silicon as
The advantages originate from the high surface-to-volume ratio, thin as a monolayer MoS2 , so the futuristic chips using MoS2 will
unique electrical, mechanical, magnetic, and optical properties. be smaller than state of the art silicon chips. Reduced electricity
TMDCs are inorganic layered materials exhibiting a variety of consumption is another advantage, along with mechanical flexibil-
electronic features ranging from semi-conductivity to supercon- ity. Moreover, the benefit of MoS2 comes from the fact that it is a
ductivity making these layered materials potentially useful in 2D material whereas silicon is 3D. MoS2 seems to be a better alter-
next-generation nanoelectronics. Materials such as MoS2 , MoSe2 , native than Si for transistors manufacturing. This consents one to
WS2 , WSe2 , h-BN, Bi2 Te3 , and Bi2 Se3 have been considered as foretell size-dependent properties and interchange collaborations
prospective materials to fabricate nanoelectronic devices due to for a broad range of semiconductors regardless of their size, shape,
their properties and ease of fabrication [6]. Recently, Late et al. composition and synthetic protocol. MoS2 is also being used in the
[7a] reported rapid synthesis and characterization of various other fabrication of solar cells and LED which leads that this material may
dichalcogenides, i.e., GaS and GaSe using a micromechanical cleav- have significant future possibilities for various field of electronics
age method and their optical properties. Moreover, single-layer and futuristic gadgets.
GaS and GaSe nanosheets represent good field effect mobility Present review summarizes the structure and synthesis of
which can be beneficial for many nanoelectronic devices [7b]. mono/few layers of MoS2 fabricated by numerous techniques
Among all TMDCs, MoS2 exhibit superior optoelectronic and cat- such as mechanical and chemical exfoliation, RF-sputtering, atomic
alytic properties compared to the conventional semiconductors. layer deposition (ALD) and chemical vapor deposition (CVD), etc.
It is vital to fabricate transistors based on these materials for Extensive theoretical aspects combined with the experimental
benchmarking essential transport characteristics toward possi- observations to provide more insights on the dielectric and topolog-
ble use in device applications. The monolayers of MoS2 have a ical behavior of MoS2 are highlighted. Additionally, broad studies
direct band-gap of ∼1.8 eV [7,8]. The successful fabrication of FETs based on potential applications of MoS2 like FET device, sensor,
with high ON/OFF ratio and NOR logic operation, phototransis- solar cells, field emission and as a catalyst for hydrogen production
tors, and gas sensors show the versatility of single layer MoS2 by of electrochemical and photoelectrochemical water splitting
[8]. have been included.
244 A.K. Singh et al. / Applied Materials Today 13 (2018) 242–270

2. Properties of transition metal dichalcogenides naturally), however can be obtained in the form of single layer
by chemical or electrochemical exfoliation from its bulk counter-
TMDCs with an MX2 stoichiometry (III-VI group, and IV-VI group part (2H-MoS2 ) [8i]. The metastable 1T has one layer per unit cell
compounds) possess similar crystal structure like layered insula- with Mo atoms in octahedral symmetric arrangement. Because of
tor (h-BN); single element materials like black phosphorus (BP), the metastable nature 1T-MoS2 , can be readily transformed to 2H
silicene and germanene; V-VI group of topological insulators (TIs) phase via intralayer atomic gliding under specific conditions [8i].
e.g. Bi2 Te3 , Sb2 Se3 and Bi2 Se3 ; transition metal oxides/hydroxides, Naturally occurring 2H-MoS2 is semiconducting while 1T-MoS2 is
such as MoO3 , V2 O5 , Ni(OH)2 and others (including metal-organic metallic in nature [8a–c]. Fig. 1(b) shows the single-layer MoS2
frames and mica etc.).8 By chemical compositions, individual layers crystal structure top and side view [7a].
of bulk TMDCs (MX2 ) is supported by a transition metal (M = Mo,
W, Ti, etc.) atomic layer sandwiched between two chalcogens (X = 2.2. Electronic band structures
S, Se, and Te, etc.) atomic layers.
Recent advancement in nanoscale characterization and device
2.1. Crystal structures fabrication have opened up new opportunities for 2-D materials
providing better insights in structure and their possible applica-
In the present review, the emphasis is on MoS2 , depending on tions. Graphene retains remarkable high charge carrier mobility
the atomic stacking configurations, the MoS2 crystal structure can at room temperature, high optical transmittance, and flexibility
be classified into 4 Polytypes: 1H, 1T, 2H and 3R according to a [1,7e]. The electronic band structure of graphene has a linear dis-
recent report [8g,h]. However, MoS2 exhibiting three structural persion near the K point [8d]. The bulk MoS2 crystal is built up
polymorphs: 2H, 3R, and 1T (Fig. 1(a)) are well reported [8i,j]. The of van der Waals bonded by S-Mo-S units. Each of these stable
structural changes directly connected to the physical properties of units consisting two hexagonal planes of S atoms sandwiching a
MoS2 . Naturally occurring bulk MoS2 possess thermodynamically hexagonal plane of Mo coordinated through ionic-covalent inter-
favored 2H phase with the S-Mo-S layers stacked in hexagonal actions with each other in a trigonal prismatic arrangement as
symmetry. On the other hand, 3R-polytype of MoS2 shares the shown in Fig. 2 [9b]. First principles, tight-binding approximations,
MoS6 with 2H in each single layer with rhombohedral symme- and many spectroscopy tools showed general electronic feature
try. The third polytype (1T phase) is metastable (not occurring of many TMDCs having similar band structures [9,10]. In general,

Fig. 1. (a) Crystal structure of MoS2 , 2H (left), 3R (center) and 1T (right) respectively. Atom color: Red, Mo; yellow, S. (Reproduced with permission from ref. 8(i) Chem 1 (2016)
699–726.) (b) Single-layer MoS2 crystal structure top and side view. (Reproduced with permission from ref. 7(a) Adv. Funct. Mater. 22 (2012) 1894–1905.)
A.K. Singh et al. / Applied Materials Today 13 (2018) 242–270 245

Fig. 2. Lattice structure of MoS2 in both: in and out-of-plane directions and clear band structure of bulk MoS2 . c1 is showing the lowest conduction band while v1 and v2
the highest split valence bands. A and B represented the direct- gap transitions, and I indicates the indirect-gap transition. Eg ís the indirect band gap for the bulk, and Eg is
the direct band gap for the monolayer. (Reproduced with permission from ref. 9(b) Phys. Rev. Lett. 2010, 105, 136805.)

Fig. 3. Band structures for bulk and monolayer MoS2 calculated from first-principles
density functional theory (DFT). The pointer indicates the fundamental bandgap Fig. 4. The variations of electrical conductivity of NbSe2 , MoS2 , and graphene as a
(direct or indirect) and the horizontal dashed lines indicate the Fermi level. The function of gate voltage at room temperature. The inset shows the typical devices
top of the valence band and bottom of the conduction band are highlighted in used for such measurements. (Reproduced with permission from ref. 11(b) Proc. Natl.
blue and green color, respectively. (Reproduced with permission from ref. 9(a) Nature Acad. Sci. USA, 2005, 102, 10451–1045.)
Nanotechnology, 2012, 699.)

2.3. Electrical and optical properties


MoX2 and WX2 compounds are semiconducting whereas NbX2 and
TaX2 are metallic [9,10]. The band structures calculated from first- The properties of TMDCs are entirely different from graphene.
principles density functional theory (DFT) for bulk and monolayer Graphene shows exceptionally high carrier mobility exceeding
MoS2 are shown in Fig. 3. The band gap transition is indirect for 106 cm2 V−1 s−1 at low temperature (2 K) and 105 cm2 V−1 s−1 at
the bulk material at the -point but gradually shifts to be direct for room temperature for devices encapsulated in BN dielectric lay-
the single layer. Due to quantum confinement effect, change in the ers [11a,b]. But, graphene does not have a bandgap so FETs made
band structure with layer number is observed, resulting in a shift from graphene cannot be switched efficiently off and having low
in hybridization between pz orbitals on S atoms and d orbitals on on/off switching ratios. As already mentioned, the bulk MoS2 pos-
Mo atoms [9c,10d]. At the same time the electronic distributions sess an indirect band gap of 1.2 eV, while single-layer MoS2 is a
are also spatially correlated to the atomic structure. semiconductor with a direct band gap of 1.8 eV. K. S. Novoselov et al.
For MoS2 , DFT calculations indicate that the energy states of first characterized electrical properties of MoS2 and compared with
the conduction-band at the K-point are because of localized d other two-dimensional materials as shown in Fig. 4 [11c]. They
orbitals on the Mo atoms, located in the middle of the S-Mo-S found the electrical conductivity of MoS2 (3 cm2 V−1 s−1 ), was much
and relatively unaffected by interlayer coupling [9a]. Conversely, lesser than graphene.
the energy states close to the -point resulting combinations of In FET structure, 2D TMDCs has been used as a semiconduct-
antibonding pz-orbitals on the S atoms and the d orbitals on Mo ing channel region, which is connected to the source and drain,
atoms holding substantial interlayer coupling effect [8e]. There- separated by a dielectric layer from the gate electrode [11c]. The
fore, as the layer numbers change, the direct excitonic states near current flowing between the electrodes of origin and drain is con-
the K-point are relatively unchanged, however the transition at the trolled by the gate voltage modulating the resistivity of the channel.
-point shift considerably from an indirect to a direct one [9b]. The Silicon is the first material that meets the industrial requirements
MoS2 undergo a similar indirect-to direct band gap transition with for performance and manufacturability in digital logic [11c]. The
decreasing layer numbers. The bandgap varies from energy range desirable properties of digital logic transistors are high current
1.1–1.9 eV [9,10]. More detailed theoretical description based on on/off ratio (the ratio of on-state to off-state conductance) for effi-
density functional theory (DFT) revealing some intriguing results cient switching, high charge-carrier mobilities for fast operation,
on the structure and properties of MoS2 are highlighted in the improved conductivity (the product of charge density and mobil-
theoretical aspect section hereafter. ity) and low off-state conductance. The on/off ratios between 104
246 A.K. Singh et al. / Applied Materials Today 13 (2018) 242–270

Fig. 5. (a) Carrier mobility calculated from first-principles DFT calculations for the electronic band structure, phonon dispersion and electron-phonon interactions of mono-
layer MoS2 as a function of temperature and. (b) carrier density, In figure (a), the gray band shows the uncertainty in calculated mobility values due to a 10% change in
computed deformation potentials associated with phonons. (Reproduced with permission from ref. 9(a) Nature Nanotechnology, 2012, 699.)

and 107 for the logic circuits are highly desirable. Two-dimensional material. The mobility of carriers is affected by the following main
MoS2 fulfilled all these criteria because of sizeable bandgap to sup- scattering mechanisms: (i) acoustic and optical phonon-scattering;
port high on/off ratios while maintaining high carrier mobilities (ii) Coulomb scattering at charged impurities; (iii) surface interface
and scalability to ever-smaller dimensions. The unique proper- phonon scattering; and (iv) roughness scattering [9a]. The scatter-
ties like excellent mobility and structural stability comparable to ing is also affacted by the thickness of the layer, device temperature,
Si make MoS2 an ideal candidate for FETs application [11d]. One effective mass and density of carrier, electronic and phonon-band
of the earliest uses of MoS2 with back-gated configurations in structure. It is well known that carrier mobility is increased by
FETs was reported in 2005, resulting high mobility values in the phonon scattering with increasing temperature. The temperature-
range 0.1–10 cm2 V−1 s−1 [11c]. Recently, Kis and coworker reported dependent electronic mobility of single-layer MoS2 calculated from
the implementation of a top-gate transistor based on monolayer first principles by Kaasbjerg et al. is presented in Fig. 5(a) [12d]. At
MoS2 [3e]. This device showed n-type conduction, with the room- higher temperatures, the optical while at lower temperatures (T <
temperature mobility of >200 cm2 V−1 s−1 , excellent on/off current 100 K) acoustic component dominates. The Coulombic scattering
ratio (∼108 ), and a subthreshold swing of 74 mV dec−1 [3e]. The in 2D TMDCs is caused by randomly oriented charged impurities
high-k dielectric HfO2 was used in the device which help to improve within the layer or on the surfaces and is the dominant scattering
the mobility MoS2 single layer. The top-gated geometry allows a at low temperatures. As a result, the mobility is also decreased, so
reduction in the voltage necessary to switch the device while per- the choice of doping in a particular device can strongly influence
mitting the integration of multiple devices on the same substrate. its performance. Fig. 5(b) shows the effect of carrier concentration
FET fabricated with MoS2 obtained from liquid exfoliation, and CVD and temperature on the carrier mobility of MoS2 .
methods show similar electrical performance [11e,12a–c]. The optical properties of MoS2 are highly intriguing. Recently,
MoS2 not only compete with conventional III–V transistors on Mak et al. investigated optical properties like absorption, PL, and
mobility values, its attractive electrical performance, relatively high photoconductivity spectroscopy on mono and few-layer MoS2 sam-
natural abundance and a high degree of electrostatic control makes ples [9b]. For PL studies, MoS2 samples were suspended and excited
it a viable for low-power electronic [9a]. The charge transport and with a CW solid-state laser with an excitation wavelength of
carriers scattering for 2D TMDCs are confined to the plane of the 532 nm. Fig. 6(a) shows the measured PL intensity under identical

Fig. 6. (a) PL spectra of mono- and bilayer MoS2 samples. The inset shows the PL QY of thin layers for N = 1–6. (b) Normalized PL spectra by the intensity of the peak of thin
layers of MoS2 for N = 1–6. Intensity for N = 4–6 is magnified, and the spectra displaced for clarity. (c) Band-gap energy of thin layers of MoS2 as a function of layers. The
dashed line shows the (indirect) band-gap energy of bulk MoS2 . (Reproduced with permission from ref. 9(b) Phys. Rev. Lett. 2010, 105, 136805.)
A.K. Singh et al. / Applied Materials Today 13 (2018) 242–270 247

excitation at 2.33 eV for a suspended monolayer and a bilayer sam- MoS2 illustrated in Fig. 8. Only two, out of four, (E1 2g and A1g )
ple. The PL quantum yield (QY) drops steadily with increasing thick- modes near 400 cm−1 are observed. The other two modes (E1g , E2 2g )
ness N = 1–6 presented in the inset of Fig. 6(a). A PL QY of the order of could not be detected either due to selection rules for scattering
10−5 to 10−6 was estimated for few-layer samples of N = 2–6, a value geometry (E1g ) or because of the limited rejection of the scattered
as high as 4 × 10−3 was observed in the limit of monolayer thick- Rayleigh radiation (E2 2g ) [13b]. Single-layer MoS2 exhibits a strong
ness [9a]. PL for bulk MoS2 is an elusive phonon-assisted process in-plane vibrational mode near ∼384 cm−1 corresponding to the
and is known to have negligible QY, similar to an indirect band- E1 2g mode. Some report claims the similar observations with dif-
gap material. The normalized PL spectra for mono- and few-layer ferent laser wavelength [13d,e]. Lee et al. reported that the E1 2g
samples in Fig. 6(b) revealed their discrete nature. The PL spec- vibration shifted toward lower wave number (redshifts), while the
trum of suspended single-layer MoS2 samples consists of a single A1g to upper wave number (blue shifts) with increasing sample
narrow feature of ∼50 meV width, centered at 1.90 eV. While few- thickness [13c,14a]. The observed upper shift of A1g peak with
layer samples exhibit multiple emission peaks (labeled A, B, and I). increasing layer number is consistent with the predicted stiffen-
Peak A coincides with the monolayer emission peak, and shifts to ing, while the behavior of the E1 2g mode doesn’t. This conflicting
the red side and broadens slightly with increasing layer numbers. result could reveal the presence of additional interlayer interac-
The thickness dependent bandgap is shown in Fig. 6(c) [9a]. tions; which also indicate that the implicit assumption of stacking
affecting the intralayer bonding is incorrect. It is monotonic as a
2.4. Vibrational properties function of film thickness (Fig. 8b). Raman mapping for the E1 2g
and A1g mode show these opposing shifts with layer thickness as
The vibrational and other physical properties of layered mate- presented in Fig. 8(c) and (d), respectively [13b]. For four or more
rials (i.e. graphene and TMDCs etc.) are completely distinct than layer system, the frequencies of both modes tend to the bulk val-
their bulk counterpart. Since vibrational properties are sensitive ues [13b]. The aforemention observations suggest that the Van der
to thickness, therefore it is necessary to study the material prop- Waals force play a trivial role on stacking of MoS2 [13b]. A detail
erties with varied thickness. Raman spectroscopy being a fast, on position of different corresponding frequencies of E1 2g , and A1g
nondestructive, high-resolution technique have been employed to peak examined using different laser lines are provided in Table 1
determine the number of layers, as well as to examine the thickness [13c].
dependent properties of various materials [12e,13a,b]. A detailed
study conducted by C. Lee et al. had nicely explained how the 2.5. Mechanical properties
weak van der Waals-like interlayer interactions affect the inter-
layer bonding, lattice vibrations and the overall properties of the MoS2 could be a promising material for next-generation flexible
system [13c]. It is revealed from the study that going from twelve electronic devices. The mechanical properties of single-layer MoS2
layers to monolayer, MoS2 exhibits exciting physical properties were first reported by Bertolazzi et al. [14a]. Bertolazzi and co-
which are absent in bulk crystal. The optical image of the differ- worker mechanically exfoliate single and bilayer MoS2 from bulk
ent MoS2 layer is shown in Fig. 7, the number of the layers was and transferred to a substrate with an array of microfabricated cir-
confirmed by AFM. A Raman spectrum for single- and few-layer cular holes as shown in Fig. 9. The Mechanical properties of the

Fig. 7. (a) Optical image of MoS2 layers deposited on the SiO2 /Si substrate. (b) AFM height indicated by dotted lines in (a). The thickness of each layer is taken along the blue
line in the AFM image and height profile is shown by a green line. (c) Schematic structure of the MoS2 bilayer. (Reproduced with permission from ref. 13(b) ACS Nano, 2010, 4,
2695–2700.)
248 A.K. Singh et al. / Applied Materials Today 13 (2018) 242–270

Fig. 8. (a) Raman spectra of thin and bulk MoS2 samples. The solid line for the 2L spectrum represents a double Voigt fit through data (circles for 2L, solid lines for the rest).
(b) Frequencies of E1 2g and A1g Raman modes (left vertical axis) and their difference (right vertical axis) as a function of layer thickness. (c, d) Spatial maps of the Raman
frequency of E1 2g (c) and A1g modes (d) for the sample in (e) Atomic displacements of the four Raman-active modes and one IR-active mode (E1u ) in the unit cell of the bulk
MoS2 crystal as viewed along the [1000] direction. (Reproduced with permission from ref. 13(b) ACS Nano, 2010, 4, 2695–2700.)

Table 1
Summary of the E1 2g and A1g peak frequencies with various laser lines “1L”, “2L”, “3L”, and “4L”, indicate monolayer, bilayer, trilayer, and four layers, respectively. A single
data represents the average value from three different samples. The measurement results with 514.5 nm line are extracted from reference [3(e)]. Ref 13(c) Adv. Funct. Mater.
2012, 22, 1385–1390.

Laser lines [nm] E1 2g peak frequency [cm−1 ] A1g peak frequency [cm−1 ]

1L 2L 3L 4L Bulk 1L 2L 3L 4L Bulk

325 384.2 382.8 382.8 382.7 382.5 404.9 405.5 406.3 407 407.8
488 384.7 383.3 383.2 382.9 383 402.8 405.5 406.5 407.4 408
514.5 [15] 384.3 383.2 382.7 382.7 382 403 404.8 405.8 406.7 407.5
532 384.7 382.5 382.4 382.4 383 402.7 404.9 405.7 406.7 407.8
632.8 385 383.8 383.8 382.9 381.5 403.8 404.8 405 406 406.6

suspended, free-standing membranes were probed with indenta- Nm−1 as shown in the left panel of Fig. 11. The pre-stress ␴0 2D
tion experiments using an atomic force microscope (AFM) with a is in the range of 0.02 to 0.1 Nm−1 . The Young’s modulus was
standard silicon cantilever (Fig. 9(c)). Bertolazzi et al. found that found to be EYoung = 270 ± 100 GPa by assuming the thickness of
the in-plane stiffness of monolayer MoS2 is 180 ± 60 Nm−1 , corre- monolayer is 0.65 nm. The Young’s modulus of bilayer MoS2 was
sponding to an effective Young’s modulus of 270 ± 100 GPa. These obtained ∼200 ± 60 GPa. The average of maximum stress for sin-
values are comparable to steel. gle and bilayer MoS2 membranes were found to be 15 ± 3 and
The force–deflection curves for mono- and bilayer MoS2 are 28 ± 8 Nm−1 , respectively (Fig. 12). The lower value of Young’s
shown in Fig. 10. Furthermore, Bertolazzi et al. extracted the modulus for bilayer MoS2 may be due to defects or interlayer slid-
pretension  0 2D and the membrane elastic modulus E2D from a ing.
least-squares fit of the experimental curves with the Eq. (1) Gomez et al. measured the mechanical properties of few layers
(5–7 layers) MoS2 similarly as disused for single and bilayer MoS2
q3 ı3 [14c] and shown in Fig. 12(a). Fig. 12(b) shows typical force vs.
F = 02D ı + E 2D (1)
r2 deformation traces (F(ı) traces measured at the center of the sus-
Calculating these parameters for a total of nine monolayers pended part of MoS2 layers), and the shape of the traces indicate
samples and averaged for the elastic modulus E2D of 180 ± 60 its thickness dependent. The thinnest sheets (5–8 layers) shows
A.K. Singh et al. / Applied Materials Today 13 (2018) 242–270 249

Fig. 9. (a) The Optical image of a single layer MoS2 flake transferred onto the pre-patterned SiO2 substrate containing an array of circular holes of diameter 550 nm. (b)
AFM image of the same monolayer MoS2 . (c) Schematic illustration of the indentation experiment. The AFM tip is placed above the center during measurements and slowly
lowered while monitoring its deflection. (Reproduced with permission from ref. 14(a) ACS Nano, 2011, 5, 9703–9709.)

films Fig. 13(a) [14c]. The histogram of the pre-tension and Young’s
modulus were obtained for 13 membranes 5–10 layers thick by fit-
ting the force vs. deformation traces to Equation 2 and shown in
Fig. 13(c) and (d).
    
4E t3 q3 Et
F= ı + (T ) ı + ı3 (2)
3(1 − 2 ) R2 R2

The strength of individual single layer MoS2 is between 6 and


11% of their Young’s modulus, and it also reflects the inherent
strength of interatomic bondings [14c]. Bertolazzi et al. also com-
pared these values to several other common engineering materials
as presented in Table 2 and found that the strength of single-layer
MoS2 was exceeded only by carbon nanotubes and graphene [14a].
MoS2 has a lower Young’s modulus and strength than graphene.
The mechanical properties of MoS2 layers indicate that MoS2 could
be suitable for integration in flexible electronic circuits.

3. Theoretical aspects
Fig. 10. The loading curves for single and bilayer MoS2 and the least-squares fit of
the experimental indentation curves to Eq. (1). The symbol × marked the fracture
point of the Membranes. (Reproduced with permission from ref. 14(a) ACS Nano, 2011, 3.1. First principles perspective of MoS2
5, 9703–9709.)
Computer simulations are emerging as a bridge between experi-
ment and theory. In particular, simulations based on first principles
the strongly nonlinear F(ı) curve, while sheets thicker than ten methods play a crucial role, because they not only help to explain
layers F(ı) traces are linear [14c]. Young’s modulus (E) and the experimental results but also can predict properties of new materi-
initial pretension (T) of suspended MoS2 nanosheets (thin layer) als or new structures with high accuracy [15a,b]. High-level density
extracted from the non-linear F(ı) curve. E = 0.35 ± 0.03 TPa and functional theory (DFT) calculations have emerged as a robust
T = 0.05 ± 0.02 Nm−1 was calculated for 8 layer from F(ı) curve as approach toward mechanistic insights into the properties of 2D
shown in Fig. 13(a). Gomez et al. also calculated E and T for thicker materials. The growth of ab-initio based physics is the outcome
250 A.K. Singh et al. / Applied Materials Today 13 (2018) 242–270

Fig. 11. Young’s modulus E2D (left) and maximum breaking stress ␴-max 2D at the central point (right) of the film for single and bilayer MoS2 flakes extracted from the
experimental data. (Reproduced with permission from ref. 14(a) ACS Nano, 2011, 5, 9703–9709.)

Fig. 12. (a) Topography AFM image of a 3–4.2-nm-thick (5–7 layers) MoS2 flake deposited on SiO2 /Si substrate pre-patterned with an array of holes 1.1 ␮m in diameter.
Inset image shows line profile acquired along the dashed line, (b) Force versus deformation curve measured at the center of the suspended part of MoS2 nanosheets with 5,
10, and 20 layers. The slope of the curve around zero deformation is plotted as a dotted line. (Reproduced with permission from ref. 14(c) Advanced Materials, 24, 6, 772–775
(2012).)

of the rapid development of computer technology. Given the crys- The most popular method for addressing the problem is the
talline atomic structures, the DFT method is capable of predicting application of density functional theory (DFT) which was devel-
the electronic, magnetic, and optoelectronic properties [15c]. One oped by Hohenberg, Kohn, and Sham and is based on the electronic
of the exciting example is the discovery of topological materials density distribution (r) rather than the many-electron wave func-
aided by computational design from first principles [15d]. The first- tion. Kohn-Sham (KS) theory is based on a basic ansatz, namely that
principles computational design approaches have worked out as a the ground state density, (r), of an interacting system is also the
powerful research tool in realizing theoretical models of various ground state density of a non-interacting system with an efficient
phases in real materials systems, outputting relevant properties potential Veff (r). The electron density is, therefore, determined from
that can be tested and confirmed directly by the experiments. a single-particle Schrodinger equation [16a]:
 1 
− ∇ 2 + Veff (r) i (r) = εi i (r) (4)
3.2. Background 2
where i (r) are the single electron KS orbitals giving the density:
A detailed description of the quantum mechanical behavior of
(r) = i |i (r)|2 , where the sum is over the occupied states. The
atoms requires explicit consideration of interactions between elec-
effective potential Veff (r) is given by where Vext (r), EH and EXC repre-
trons and nuclei. The Born–Oppenheimer approximation is the
sent the Coulomb interaction between electrons and ions, Coulomb
initial approximation used to solve the Hamiltonian by because the
interaction among the electrons and exchange-correlation interac-
nuclei cannot move as much as electron due to its massive mass. As
tion, respectively. The effective potential depends on the electron
a result, the motions of the two subsystems (electrons and nuclei)
density that is determined by the Kohn-Sham orbitals, which in
can be uncoupled. Thus the Hamiltonian for given N electrons mov-
turns depends on the effective potential. Thus the problem of solv-
ing in the external potential (Vext ) of fixed nuclei in atomic units
ing the Kohn-Sham equation is iterative with the requirement
 = m = |e| = 1 becomes [16a]:
that the effective potential and the resulting charge density are
self-consistent. The validity of effective implementation of the
1 2  1 1
i
H=− ∇i + Vext (ri ) + (3) Kohn-Sham method lies in finding a good approximation for EXC
2 2 rij and hence Veff [15a,16a]. The simplest approach is the local density
i i ij
approximation (LDA) where the contribution of each volume ele-
An analytic solution of this many-body Schrodinger equation ment to the total exchange-correlation energy is taken to be that
is impossible to obtain because of 3N degree of freedom for each of an element of a homogeneous electron gas with the density cor-
electron. responding to that point. One strategy to improve the LDA is to
A.K. Singh et al. / Applied Materials Today 13 (2018) 242–270 251

Fig. 13. (a) Force versus deformation curve for eight layers thick suspended MoS2 over a hole 1.1 ␮m in diameter. The dotted black trace is the fit to Eq. (2), employed to
obtain the Young modulus E = 0.35 ± 0.02 Tpa and the pre-tension T = 0.05 ± 0.02 N m−1 . (b) Force–volume measurement is showing a color map of the compliance (inset) and
its radially-averaged profile of the sheet. (c) Histogram of the initial pre-tension obtained with the help expression (2) for 13 sheets 5 to 10 layers thick MoS2 . (d) Histogram
of Young’s modulus obtained from fitting F(d) curves to expression (2) for the same 13 sheets plotted in the panel. (Reproduced with permission from ref. 14(c) Advanced
Materials, 24, 6, 772–775 (2012).)

include the gradient of the charge density in exchange-correlation different basis sets many simulation packages like VASP, SIESTA,
functional, which is given in generalized gradient approximation CASTEP, etc. have been developed to solve Kohn-Sham equations.
(GGA) [15a,16a]. Usually, valence electrons determine the physical properties of
Also, for a crystalline solid, the Bloch theorem can be used a solid. Therefore, the pseudopotential approximation can be
to impose boundary conditions for given external potential and introduced, which removes the core electrons and replaces the
electron wave function, resulting in a set of coupled second-order strong nuclear potential with a weaker pseudopotential [15a,b].
differential equation, each describing an electron moving in an
average potential due to the other electrons [16b]. This set of 3.3. Electronic structure of MoS2 as a function of layers thickness
the equation can be for solid-state systems can be solved using
localized atomic orbitals or plane-wave expansion of the wave Electronic structure of bulk MoS2 finds it as an indirect-band-
function and k-point sampling in reciprocal space. Depending upon gap-semiconductor [17a–d]. DFT calculations have shown that
252 A.K. Singh et al. / Applied Materials Today 13 (2018) 242–270

Table 2
Comparison of Young’s moduli and breaking strengths for several engineering materials including monolayer MoS2 [4d,5d,8a,e,9d-e,10a-c].

Material Young’s modulus Breaking strength Breaking strength (%)


Eyoung (GPa)  max eff (GPa) Young’s modulus

Stainless steelASTM-A514 205 0.9 0.4


Molybdenum 329 0.5–1.2 0.15–0.36
Polyimide 2.5 0.231 9
PDMS 0.3–0.87 2.24 2.5
Kevlar 49 112 3 2.6
Monolayer MoS2 270 16–30 6–11
Bulk MoS2 238
WS2 nanotubes 152 3.7–16.3 2.4–10
Carbon nanotubes 1000 11–63 1.1–6.3
Graphene 1000 130 13

Fig. 14. Electronic band structure of bulk MoS2 , its mono, and multilayers. The top of the valence band (blue) and bottom of the conduction band (red) are highlighted. The
arrow indicates the smallest value of the indirect or direct band gap. The Fermi level is set at 0 eV. (Reproduced with permission from ref. 17, Eur. Phys. J. B., 2012, 85, 186.)

when MoS2 is thinned to a single layer, an indirect to direct band 3.4. Dielectric response of MoS2 : electron energy loss
gap transition occurs [17a–c]. While downsizing the layers from spectroscopy
bulk to the single layer limit, the indirect bandgap increases and
becomes more substantial for the monolayers that the materi- Electron energy loss spectroscopy has been suggested as a pow-
als converted to direct-band-gap semiconductor (Fig. 14), thereby, erful technique to identify the number of layers in commercially
causing a gradual blue shift (1.14 eV) in indirect band gap energy produced large and uniform 2D sheets [18a]. Electron energy loss
[17c]. At the same time, magnitudes of the blue shift in the direct spectra (EELS) can also give information about the single electron
bandgap energies at the ‘K’ point is found to be 0.19 eV. Therefore, interband transitions which can be identified in the lower energy
the change in the indirect bandgap energy is significantly enhanced side from the collective plasma oscillations [18b]. Theoretical EELS
than that of the direct band gap energy. The blue shift in the band can be calculated from a dielectric function using the expression:
gap energies is maximum when one goes from bilayer to the mono-
layer of MoS2 . The conduction band states at ‘K’ point are mainly −1 ε2 (ω)
Im = (5)
due to the localized’ orbitals of the Mo atoms, sandwiched between ε(ω) ε1 (ω) + ε22 (ω)
2

two “S” layers (“S-Mo-S”) and relatively unaffected by interlayer


coupling. However, the states near the ‘’ point are due to the com- where ε1 , ε2 are real and imaginary parts of the dielectric func-
binations of antibonding pz orbitals of S atoms and the d orbitals of tion. Note that imaginary part of the dielectric function (ε2 ) can be
Mo atoms and have a strong interlayer coupling effect [17e]. There- obtained by first-order time-dependent perturbation theory using
fore, as the layer number changes, the direct excitonic states near ground state DFT eigenvalues and eigenfunctions [18c]. The real
the k-point are relatively unchanged, but the transition at the ‘’ component of the dielectric function (ε1 ) can easily be obtained
point shifts significantly from indirect one to a larger, direct one. from ε2 using Kramers–Kronig relation. The calculated EELS for
Therefore, the change in the band structure with layer number is MoS2 consists of two prominent resonance features for in-plane
due to quantum confinement and the resulting change in hybridiza- polarization (E⊥c) that lies (i) below 10 eV which is due to ␲-
tion between the pz orbitals on S atoms and d orbitals on Mo atoms. plasmons (␲-␲* transitions) and (ii) above 10 eV which is due to
This, indirect-to-direct band gap transition in MoS2 has opened up ␲+␴ plasmons (␲-␴* and ␴-␴* transitions) [Fig. 15(a)], whereas for
a way for potential applications in photonics, optoelectronics, and out-of-plane polarization (E || c) only one resonance feature above
sensing [17f]. 10 eV due to ␲-␴* and ␴-␴* transitions, is prominent [18d]. In MoS2
A.K. Singh et al. / Applied Materials Today 13 (2018) 242–270 253

Fig. 15. (a) EELS of MoS2 for E⊥c and E||c with some layers. (b) The variation in the εs concerning some layers for both E⊥c and E||c. (c) The imaginary (ε2 ) of the dielectric
function for bulk and monolayer MoS2 for both E⊥c and E||c. (Reproduced with permission from ref. [18(d)], Physica B., 2012, 407, 4627–4634.)

with 18 valence electrons, the strong covalent bonding between the results into virtually no change in the interband transition peaks in
atoms of Mo and S is made up of 12 electrons per molecule having ε2 . However, the intensity of the interband transition peaks in the ε2
mixed s, p and d orbitals which results into the  valence band and decreases significantly which is attributed to reduced in the num-
the remaining six electrons are responsible for the ␲-plasmon band ber of bands (Fig. 14) for transition when the thickness of MoS2 is
[18b]. Among the six electrons in the ␲-plasmon band, two occupy diminished to the monolayer.
the dz 2 orbitals of the metal atom.
It is found that the ␲+␴ plasmons peak for both perpendicu- 3.5. A topological aspect of MoS2
lar and parallel polarization remains broad. The loss spectra shift
toward lower energies even by the variation of slab thickness from The discoveries of novel materials often drive the progress
bilayer to monolayer [18d,e]. -plasmon peaks shift marginally, in materials science. In this regard materials presenting unique
but prominent redshift can be seen in ␲+␴ plasmon peaks for both quantum mechanical properties are of particular importance. The
perpendicular and parallel polarization as one goes from bulk to quantum Hall (QH) state discovered in 1980 [19a] is the first exam-
monolayer limit. It emerges that ongoing from bulk to the mono- ple of a quantum state that does not have spontaneously broken
layer, the average concentration of electrons decreases which result symmetry, which was topologically distinct from all previously
into a considerable enhancement in the effective mass of electrons known states of matter. Later on, in 1982, the concept of topological
and hence a lower value of ωp for monolayers, which is consistent description was used to explain the quantization of the Hall conduc-
Ne2
with the equation: ωp2 = ε0 meff , where N, ε0 , e, and meff are number tance in two dimensional (2D) electron gases, which gives rise to a
density of electrons, permittivthe ity of the free space, charge onthe critical characteristic is known as Thouless-Kohmoto-Nightingale-
the electrons and effective mass of the electrons, respectively. Note Nijs (TKNN) number that is topologically invariant [19b]. To realize
that experimental EELS of MoS2 also show keen sensitivity to the quantum Hall effect (QHE) of materials, a strong external magnetic
number of layers [18f]. Therefore, EELS can be used as a technique field is required. Consequently, an alternative theoretical model
to identify the number of layers in 2D sheets of MoS2 . was proposed by Haldane for 2D honeycomb-lattice [19c]. The Hal-
Static dielectric constant (εs ) is an essential parameter for device dane model was further modified by Kane and Mele [19d] which
modeling which can be obtained from a real part of dielectric func- reveals that the intrinsic spin-orbit coupling (SOC) induces novel
tion at zero frequency. A pronounced red shift in the value of εs gapless edge states that counter-propagate at the boundaries with
has been reported for both perpendicular and parallel polariza- opposite spins, and are protected by time-reversal symmetry (TRS)
tion when MoS2 is thinned to monolayer limit [Fig. 15(b)]. The [19e], resulting in a vanishing charge Hall conductance but nonzero
redshift in the values of εs for both the polarization is consistent quantum spin Hall (QSH) conductance. This new phenomenon was
with the corresponding blue shift in the band gap energies as going termed as the QSH effect with the electronic spin as the quantum
from bulk to monolayer limit (Fig. 14) which can be verified from number and characterized by the Z2 topological order [19f]. The
the two band model proposed by Penn for semiconductors [18g]: successful predictions of new topological insulator (TI) materials
2 are perhaps the most spectacular triumph of the modern state-
ωp
εs ≈ 1 + Eg
, where ωp is plasmonic frequency, and is the band
of-the-art first-principles theory that has emerged as a vital tool
gap of given the semiconductor. The imaginary part of the dielec- to enable materials discovery by designing unknown materials as
tric function (ε2 ) that reveals the interband transitions [17d] is well as unexplored properties of existing materials that are sub-
highly anisotropic in low energy range and become isotropic in high sequently confirmed by experiments. To date, most TIs are first
energy range [Fig. 15(c)]. ε2 for MoS2 as a function of layer number theoretically predicted or computationally developed, followed by
remains nearly unchanged regarding peak shift which is under- experimental confirmation [15d,19g]. Topological insulators (TIs)
stood from the corresponding electronic band structures, where materials remain a topic of hot discussion among the researchers in
bands above and below the Fermi level resides almost constant that the recent time. The 2016 noble prize in Physics has been awarded
254 A.K. Singh et al. / Applied Materials Today 13 (2018) 242–270

for the “theoretical discoveries of topological phase transitions and Therefore, using above criterion,  = 1 at a particular TRIM point
topological phases of matter.” indicates the non-trivial TI or quantum spin Hall (QSH) state,
whereas if the computation with including spin-orbit coupling
3.6. Topological Z2 invariant (SOC) effect yield  = 0 or trivial state, then there is a band inversion
at that particular TRIM point which is accompanied by a change in
The more general mechanism for generating TIs is through an the parity of two states involved and thus changes the value of 
electronic band inversion, where the usual ordering of the conduc- [20d].
tion and valence band with different parities is ‘inverted’ by strong
SOC effect [20a], thereby, allowing the presence of 2D TI states in 3.7. 1T’-MoS2 as topological insulators
a much larger variety of materials. The topological order of elec-
tronic band structure, the, i.e., the Z2 invariant is an important Monolayer MoS2 possess a variety of polymorphs such as 1H,
parameter to describe the unique phase of matter [20b]. Among 1T and 1T’ [Fig. 16(a)]. The most studied 1H structure is a triatomic
the various computation of Z2 topological invariants, the Fu-Kane layer sandwich of Bernal (ABA) stacking with P6−m2 space group,
criterion [20c] is exceptionally well suited for analyzing the band whereas, the three atomic planes in the 1T structure form rhombo-
structures of 2D crystals. The Fu-Kane approach connects the Z2 hedral (ABC) stacking with P3−m2 space group. It has been shown
invariants to the matrix elements of Bloch wave functions at time- that the 1T structure of MoS2 is unstable in free-standing condition
reversal invariant momentum (TRIM) points in the Brillouin zone and undergoes a spontaneous lattice distortion in the x-direction to
(BZ). Note that there are four TRIM points in 2D BZ and time- form a period-doubling 2×1 distorted structure, i.e., the 1T’ struc-
reversal symmetry (TRS) yields one unique Z2 invariant, , which ture [21a], consisting of 1D zigzag chains along the y-direction
is given by: [Fig. 16(a)]. Unlike 1T-counterpart of MoS2 , 1T’-MoS2 is a semi-

n conductor with a fundamental gap (Eg ) of about 0.08 eV, located at
(−1)v = ıi (6) ‘’ point [Fig. 16(b)].
i=1
The conduction and valence bands undergo a band inversion
with a sizeable inverted gap (2ı) of about 0.6 eV at ‘’ point of 2D
where n = 4 for 2D lattice and ıi denotes the Pfaffian of a unitary Brillouin zone [Fig. 16(b) & (c)]. Note that 1T’ structure has inver-
matrix over TRIM points in BZ [20c]. In a crystal with inversion sym- sion symmetry with Z2 = 1 indicating that 2D 1T’-MoS2 possess QSH
metry, Bloch wave functions are also eigenfunctions of the parity insulator phase [21b]. Looking at the orbitals character of inverted
operator with eigenvalues m = ±1, and the formula for ıi simplifies band structure near ‘,’ it was found that the valence band mainly
as: consists of Mo d-orbitals whereas in conduction band S p-orbitals
 dominate [Fig. 16(b)]. The unexpected band inversion is attributed
ıi = m (7)
to the doubling of the Mo chain in the 1T structure, which low-
m
ers the Mo d-orbitals below p-orbitals of S-atom around ‘’ point.
where the product is over the pair, the parities of occupied Kramer’s In the absence of spin-orbit coupling (SOC), this band inversion
doublet resulting from time-reversal symmetry at the TRIM points leads to the appearance of two Dirac cones around ‘’, whereas SOC
without multiplying the corresponding time-reversal partners. opens up a fundamental gap of 0.08 eV at the Dirac points [inset of

Fig. 16. (a) Atomistic structure of three polymorphs of the MoS2 monolayer. M stands for metal (Mo), and X stands for chalcogen (S). 1H-MoS2 and 1T-MoS2 from ABA and
ABC stacking, respectively. 1T -MoS2 is distorted 1T-MoS2 structure where the distorted Mo atoms form 1D zigzag chains indicated by the dashed blue line. Red rectangles
show the unit cell. (b) Calculated electronic band structure of 1T’-MoS2 . Eg is a fundamental gap, and 2ı is an inverted gap. The inset compares band structures with (red
dashed line) and without (solid black line) spin-orbit coupling. (c) Brillouin zone of 1T’-MoS2 . Black dots marked four time-reversal invariant moments and labeled as , X,
Y, and R. Red dots mark the locations of the fundamental gap. (d) Topological phase diagram of 1T -MoS2 as a function of the vertical electric field. The critical field strength
is ±0.142 V/Å, marked by two green dots. (Reproduced with permission from ref. [21(b)], Science, 2014 346, 1344.)
A.K. Singh et al. / Applied Materials Today 13 (2018) 242–270 255

Fig. 17. (a) Top and side view of the crystal structures of monolayer 1S-MoS2 . (b) 2D Brillouin zone of monolayer 1S-MoS2 (c) Phonon band dispersion of 1S-MoS2 . (d)
Electronic band structures of monolayer 1S-MoS2 with and without spin-orbit coupling (SOC). The Fermi level is set to zero. (e) The evolution of orbitals-resolved band
structures of 1S-MoS2 under SOC effect. (f) Band structure of 1S-MoS2 around the Fermi level in 2D k-space with energy as the third dimension. (Reproduced with permission
from ref. [22(c)], Phys. Rev. B., 2015, 92, 085427.)

Fig. 16(b)], that results in QSH transitions by the similar mechanism conduction bands occur [22c]. Therefore, SOC does not induce any
as the Kane-Mele model for graphene [19d]. band inversion between the valence and conduction bands at the
The inverted bands between p orbitals of S and d orbitals of TRIM. The only effect of SOC is to break the degeneracy at the touch-
Mo located on well-separated atomic planes offer a simple mecha- ing point and to open an insulating gap. This is different from some
nism to control topological electronic properties by perpendicular well-known TIs, such as Bi2 Se3 [23a], where the SOC is responsible
electric field [21b], which is highly desirable for devices based on for the band inversion. Such characteristics of ‘intrinsic nontrivial
van der Waals heterostructures. The vertical electric field induces a band order’ have also been found in several other systems, such
topological phase transition [Fig. 16(d)] by breaking inversion sym- as the chemically modified Bi and Sn honeycomb lattices [23b,c].
metry in 1T’-MoS2 that introduces a substantial Rashba splitting of Also, a porous allotrope of monolayer MoS2 (g-MoS2 ) consisting of
the doubly degenerate bands near the fundamental gap. On increas- the square-hexagonal structure has been predicted to be QSH insu-
ing the perpendicular field, Eg first decreases to zero at a critical field lator [23d] and the calculated nontrivial gap as large as 109 meV
strength of ±0.142 V/Å and then reopens on further increase in the is higher than 1T’-MoS2 and 1S-MoS2 . In contrast to 1T’- and 1S-
magnitude of the field. This phenomenon offers the possibility to phase of MoS2 , the QSH effect in g-MoS2 originates from the pure
control on/off and charge/spin conductance of helical edge states d-d band inversion. Such new polymorph dramatically enriches the
by the external field in QSH-based devices [21c-g]. QSH insulator family based on MoS2 .
Another allotrope of MoS2 (1S-MoS2) possesses square-octagon
ring structure with D4h symmetry [22a-d] which can be viewed 4. Growth methods
as a three-layer stacking of Mo and S atoms, wherein Mo atoms
are sandwiched between layers of S atoms [Fig. 17(a)]. The equi- 4.1. Mechanical exfoliation
librium lattice constants are calculated to be 6.336 Å [22c]. The
inversion symmetry also exists in 1S structure. Note that four- Mechanical exfoliation (ME) is simply “peeling” or “cleaving”
and eight-membered rings structures found at grain boundaries the bulk crystals with the help of an adhesive tape or ‘rubbing’ bulk
of 1H-MoS2 [22e]. The calculated phonon dispersion of 1S-MoS2 crystals against a solid surface, is categorized as a top-down tech-
possesses positive frequencies of all phonon branches in the whole nique, has been widely used for the fabrication of single- and/or
Brillouin zone indicating the dynamic stability of 1S-phase of MoS2 few-layer systems from their bulk counterpart [1d,11b]. As dis-
[Fig. 17(b) & (c)]. Electronic band structure of 1S-MoS2 without con- cussed, the atoms in 2D TMDCs are covalently connected to their
sidering SOC, shows valance band maximum (VBM) and conduction neighboring layers, via van der Waals bonding along the third axis.
band minimum (CBM) touching at the same point indicating it The weak interlayer van der Waals forces enable the facile exfo-
as gapless semiconductors [22e]. VBM and CBM are mainly con- liation of the TMDCs. ME has been successfully applied for the
tributed by the dx2-y2 and dz2 orbitals of the Mo atoms, respectively fabrication of few layers of MoS2 (thickness ∼3–10 nm) [10d-g].
[Fig. 17(d)]. Besides, this the CBM + 1 at the ‘’ point is mainly con- The Scotch tape, chemical and mechanical exfoliation methodolo-
tributed by dx2-y2 orbitals. Electronic band dispersion of 1S-MoS2 in gies are the most common techniques to produce and disperse
two-dimensional k space around the touching point at Fermi level TMDC layers in polar solvents. This technique produces a very clean,
confirms the touching of VBM and CBM at same point [Fig. 17(e)]. high quality samples [10h,i]. Optical and AFM images of single layer
By turning SOC on, imposes a band exchange between CBM and MoS2 is shown in Fig. 18 and Fig. 19 respectively [4e,14c]. However,
CBM +1 at the ‘’ point that results into a SOC-induced direct band large scale production is not possible adopting this methodology,
gap of 24 meV in 1S-MoS2 [Fig. 17(d) & (e)]. so limited to prepare samples for research purpose only.
The topological Z2 invariant is calculated to be Z2 = 1 indicating
1S-MoS2 monolayer as nontrivial 2D TIs. Considering the sizable 4.2. Chemical vapor deposition
nontrivial band gaps, the QSH effect can be readily realized at room
temperature. By carefully examining the parities of the low-energy Chemical vapor deposition (CVD) is another common and
bands, it is found that no parity exchange between the valence and extensively used techniques to synthesize and understand the
256 A.K. Singh et al. / Applied Materials Today 13 (2018) 242–270

Fig. 18. Single-layer of MoS2 sheet on 300 nm SiO2 /Si substrate obtained using micromechanical cleavage method. (a) An optical image, (b) Raman image, (c) AFM image,
and (d) corresponding AFM height profile respectively. (Reproduced with permission from ref. 4(e) ACS Nano, 4879, 2013.)

where x = 1 once an intermediate phase formed. However, a system-


atic study describing detail on growth parameters are still missing
this, and the fabrication of large area MoS2 is still challenging [25c].
Recently, Henz’s and Ajayan’s group separately reported CVD-
based procedure to fabricate large-area, highly crystalline MoS2
layers by sulphurizing vapor-phase MoO3 . These experiments show
the crucial role of MoO3 during the growth of MoS2 which was
controlled by the diffusion of vapor-phase MoO3. Several growth
stages from random nucleation to small triangular domains and
their evolution to the continuous film were systematically studied.
The method is very efficient and promises large-area continuous
films of single-layered MoS2 . Their findings were in contrast to
graphene growth on Cu foil by CVD method, and with complicated
nucleation process on bare SiO2 substrates [25d,e]. These obser-
Fig. 19. Atomic force microscope image of the single layer of MoS2 on a silicon
vations strengthen the argument made by Li at al. on CVD-grown
substrate. (Reproduced with permission from ref. 14(c). Nature Nanotechnology, 2011,
6, 147–150.) MoS2 and interpreting the importance of control nucleation which
is essential for the growth of large-area MoS2 monolayers [25c].
The fundamental requirements of this approach on the edge-based
microscopic growth mechanism of 2D TMDCs. Traditionally, the
nucleation growth bear a resemblance to the observations and the-
sulphurization of the MoO3 (Molybdenum (VI) oxide) powder to
oretical predictions for the growth of some other layered materials
grow monolayer MoS2 has been accepted and widely studied
[25d,e].
[24,25a–c]. Benefited from the moderate melting and evapora-
tion temperatures, Li et al. successfully demonstrated that at
favorable growth conditions, the MoO3 powder is a suitable pre-
4.2.1. Vapor phase growth
cursor for MoS2 thin films growth using CVD technique [24,25a-c]
For the fabrications of large-area MoS2 thin films, vapor phase
(Fig. 20) [25c]. In another experiment, Li et al. describe growth
deposition stands as an appealing and versatile strategy. Sul-
mechanism of CVD grown MoS2 layers, assisted by seeding
phurization of solid MoO3 films have involved many researchers
the substrate with graphene-like species but the study lacks
and exposed a straightforward technique for the synthesis MoS2
whole grains and grain boundary characterization [25d]. Later
[26a,b]. However, the quality and the small grain size (∼20 nm)
on, Li et al. concluded that the pretreatment of substrates with
of thus grown samples is not admirable and possess low carrier
various aromatic molecules including reduced graphene oxide,
mobility (0.004–0.04 cm−2 V−1 s−1 ) in comparison to mechanically-
perylene-3,4,9,10-tetracarboxylic acid tetra-potassium salt (TPAS)
exfoliated samples (01–10 cm−2 V−1 s−1 ) [26b]. In order to fabricate
and perylene-3,4,9,10-tetracarboxylic dianhydride (PTCDA) before
high-quality, uniform and large-area TMDCs with good electri-
deposition helps in subsequent MoS2 growth [16c]. The reaction
cal performance, efforts have been devoted to the thermolysis
mechanism for the growth of MoS2 layer was proposed to be:
the precursor containing Mo and S atoms. Li et al. [26b] reported
MoO3 + x/2S → MoO3 − x + x/2SO2 and a two-step thermolysis process of ammonium thiomolybdate
resulting large-area, highly crystalline MoS2 thin films on divers
MoO3 − x + (7 − x)/2S → MoS2 + (3 − x)/2SO2 insulating substrates (Fig. 21). During the thermolysis the ammo-
nium thiomolybdate precursor dissociated into MoS2 , NH3 and S
A.K. Singh et al. / Applied Materials Today 13 (2018) 242–270 257

Fig. 20. (a) Schematic illustration of the MoS2 layers by MoO3 sulfurization. A layer of MoO3 (3.6 nm) was thermally evaporated on the sapphire substrate. The MoO3 was
then converted to a MoS2 by a two-step thermal process. (b) MoS2 layer is grown on a sapphire wafer. (c and d) Surface topographic images obtained by AFM. (e) A selected
cross-sectional height profile is showing the thickness of the MoS2 layer. (Reproduced with permission from ref. 25(b). Nanoscale, 2012, 4, 6637.)

Fig. 21. (a) Schematic illustration for the Synthesis of MoS2 thin layers on insulating substrates by two-step thermolysis process. (b) Raman spectra for the bi and tri-layer
MoS2 grown on sapphire substrates. (c) Energies of the two characteristic Raman peaks for the micromechanically exfoliated MoS2 films with different layers. (d) The PL
intensity of the tri-layer MoS2 thin films prepared in (Ar+S) is stronger than those developed in pure Ar. (Reproduced with permission from ref. 26(b), Nano Letters 78 (3),
1538–1544 (2012).)

at/above 800 ◦ C [26b]. The multistep transformation reaction from multilayer) produced by mechanical exfoliation, are unam-
(NH4 )2 MoS4 into MoS2 involves fairly complex chemistry. The tran- biguously less which is disadvantageous for their commercial
sistor devices fabricated with MoS2 thin layers using a bottom gate utilization. To obtain large quantities of mono- or few-layer
geometry exhibit n-type behavior with a significantly improved TMDCs nanosheets, a solution processing strategy is strongly
on/off current ratio [26b]. required. Recently, Coleman et al. proposed liquid-phase exfolia-
tion of bulk MoS2 powders in an organic solvent with the help of
4.3. Liquid-phase preparations ultra-sonication and demonstrated the simple fabrication of the
exfoliated material [18f]. N-methylpyrrolidone (NMP) is used as
The bulk TMDC crystals exfoliated into mono- or few-layer most popular solvent, which is universal to exfoliation of most
nanosheets found to meet their full range of potentials. How- TMDCs [18f]. This method gives thin sheets of MoS2 with thick-
ever, the yield of exfoliated TMDCs materials (i.e., single, bi- and nesses around 4–10 nm, corresponding to 6–18 layers. However,
258 A.K. Singh et al. / Applied Materials Today 13 (2018) 242–270

Fig. 22. (a) SEM images of MoS2 powder. The Inset shows higher magnification image of the layered structure. The scale bar is 400 nm. (b) Photograph of a standard chemically
exfoliated MoS2 suspension in water. (c) AFM image of individual exfoliated MoS2 sheets after annealing at 300 ◦ C. Height profile along the red line is overlaid on the picture.
(d) TEM image of an as-deposited MoS2 sheet on a holey carbon grid. (e) SAED pattern from the MoS2 layer in (d) showing the hexagonal symmetry of the MoS2 structure.
(f) HAADF STEM image of monolayer MoS2 annealed at 300 ◦ C in Ar. Inset is a blow-up image showing only Mo (green dot) and S (orange dot) atoms and their honeycomb
arrangement. The scale bar is 0.5 nm. The white blur on the surface of the sheet is possibly due to carbon residues from the intercalation and exfoliation processes. (Reproduced
with permission from ref. 28(a) Nano Lett. 2011, 11, 5111–5116.)

monolayers were difficult to produce by this technique. After investigation confirmed the growth of continuous MoS2 film. The
the success of Colemen’s pioneer work, LPE was also successfully process can be easily controlled by controlling the reaction param-
applied to synthesize 2D material phosphorene using benzonitrile eter to grow bilayer as well as few layers MoS2 films (Fig. 23). The
as liquid media under inert atmosphere [26c,f]. On the other hand, Raman and photoluminescence (PL) spectra confirmed the crys-
Morrison and co-workers reported that Li-intercalated MoS2 could talline quality of the MoS2 films. The mobility was evaluated to
be exfoliated into monolayers via forced hydration, yielding a sta- be ∼29 cm2 V−1 s−1 with the current on/off ratio ∼104 for bilayer
ble colloidal suspension [26d]. This method provides a flexible way and ∼173–181 cm2 V−1 s−1 for the few-layer MoS2 system. Hus-
toward the assembly of MoS2 sheets into thin films [27a-c] that can sain et al. concluded that the higher mobility behavior in their
be utilized for photovoltaic applications [27d]. Though, exfoliation study could be attributed to low charged impurities of the film
of MoS2 by Li-intercalation changes the fundamental properties of along with the dielectric screening effect by an interfacial MoOx Siy
MoS2 [27e]. However, the disadvantage of this method is it requires layer. The authors further argued that RF sputtering followed by
a long time (days) yielding smaller size 2D nanosheets. Fig. 22 post-deposition annealing opens up the new possibilities of mass
shows detailed characterizations of the MoS2 layers obtained by production of large area MoS2 films [28b].
liquid-phase exfoliation of bulk MoS2 powders in an organic sol- In a similar experiment, Hussain et al. discussed the efficient
vent [28a]. This method is not suitable for the production of high layer-controlled, continuous and large-area MoS2 growth onto
quality materials for their potential application in optical, elec- SiO2 /Si substrate by RF sputtering followed sulfurization adopting
trical, and electrochemical measurements. However, the products some modification to their earlier experiments [28c]. As compared
obtained from this method can be used for catalysis, supercapac- to their previous study, the number of MoS2 layers here can easily
itors and sensors etc. where the compromised material quality is be controlled by adjusting initial sputtering time. The MoS2 transis-
not an issue [27f-g]. Laser ablation of bulk crystals is an example of tors fabricated on thus grown MoS2 films exhibited high mobility
a less common top-down method [10i]. values ∼21 cm2 V−1 s−1 (bilayer) and ∼25 cm2 V−1 s−1 (trilayer), with
significantly high on/off ratios in the range of ∼107 (bilayer) and
4.4. MoS2 film fabricated by RF sputtering 104–105 (trilayer), respectively [28c].

Recently, Hussain et al. reported a simple and scalable approach 4.5. Atomic layer deposition (ALD)
for the synthesis of bi- and few layers MoS2 films via RF sputter-
ing followed by the post-deposition annealing method. In their Atomic layer deposition (ALD) is a thin film deposition method
experiment, MoS2 films were fabricated using a MoS2 target in based on self-limiting mechanism. ALD can control film thickness
the sputtering system. To improve the crystalline quality of as- at an atomic scale precision while forming highly uniform and
sputtered samples, thin films were subjected to post-deposition conformal thin films on the large-area. Recent advances in the
annealing at 700 ◦ C in Sulfur and Argon environment [28b]. The synthesis of 2D TMDCs have shown that atomic layer deposition
A.K. Singh et al. / Applied Materials Today 13 (2018) 242–270 259

Fig. 23. (a) Schematic illustration of the RF sputtering technique to prepare MoS2 thin films. Annealing at elaborated temperature was performed to enhance the crystalline
quality of sputter grown MoS2 films under Ar/S environment. (b-d) represents optical images of MoS2 films grown on SiO2 /Si substrate for (b) 1 min; (c) 3 min, and (d) 5 mins
sputtering. (Reproduced with permission from ref. 28(b) Scientific Reports, 2016, 6, 30791.)

Fig. 24. Schematic illustrations of the MoS2 film fabricated by atomic layer deposition method. The optical images show (a) 10 and (b) 50 cycles of the ALD MoS2 film grown
on a 2-inch sapphire (001) substrate. (Reproduced with permission from ref. 28(d) Nanoscale, 2014, 6, 10584.)

(ALD) of the metal oxide and subsequent sulfurization could offer a reactions of MoCl5 and H2 S precursors as shown in Fig. 24. Due
method to synthesize large-area MoS2 with excellent layer control to the self-limiting reactions of the vapors, the number of layers
over the entire substrate. In the recent efforts by Loh et al. reported can be precisely controlled by adjusting the number of deposition
on the growth of controlled single to multilayer MoS2 film by using cycles. Post-deposition annealing was introduced to improve the
the ALD technique at 300 ◦ C on a sapphire wafer [28d]. In their crystallinity as shown in Fig. 25, which was evident from the
process, ALD provides precise control over MoS2 film thickness presence of triangle-shaped crystals exhibiting the strong photo-
due to the pulsed introduction of the reactants and self-limiting luminescence in the visible range [28d]. However, ALD has some
260 A.K. Singh et al. / Applied Materials Today 13 (2018) 242–270

Fig. 25. (a) TEM image of single and few-layer MoS2 film grown on sapphire (001) substrate by ALD technique. (b) The TEM image of the MoS2 film revealing high crystallinity
of the film produced by ALD with the hexagonal lattice structure (inset image). (c) Lattice spacing (d) The line profile depicting the spacing between layer (c). Cross-sectional
TEM view of (e) 50 cycles and (f) 10 cycles ALD MoS2 film on the sapphire substrate. (Reproduced with permission from ref. 28(d). Nanoscale, 2014, 6, 10584.)

inherent challenges like growing large area oxide films with sub co-workers for the first time reported a top-gate transistor based on
1 nm nucleation, and the necessary steps are still to be unexplored. monolayer MoS2 . The device showed excellent on/off current ratio
A very recent study on the synthesis of the MoS2 film, Grossman (∼108 ), n-type conduction, the room-temperature mobility of >200
et al. [28e] reported, control atomic layer deposition of molyb- cm2 V−1 s–1 and a subthreshold swing of 74 mV per decade [14d].
denum oxide nucleation followed by sulfurization to produces The cross-sectional view of MoS2 transistors and its performance
large-area MoS2 monolayers. In their work, Grossman et al. demon- is shown in Fig. 27. Later-on Walia et al. also fabricated transis-
strated that the necessary process required achieving sub 1 nm tors based on MoS2 with aluminum, tungsten, gold, and platinum
nucleation domains formed by oxide deposition. Subsequently, fab- contacts [29b]. From their work, it was evident that lower work
rication of the MoS2 films was performed by the sulfurization of functions of the contact metals lead to a smaller Schottky barrier
oxides deposited. Large-area, uniform MoS2 films were achieved by size and thus higher charge carrier injection through the contacts.
optimizing the effects and surface treatments on the ALD nucleated The study indicates that choosing a suitable metal contact is crucial
Mo oxide and the post sulfurization process [28e]. in tuning the barrier height at the interface of the metal semicon-
For better understandings of the controlled oxide deposition, ductor. In another study, Kang et al. presented high-performance
film chemistry analysis during sulfurization, various temperature MoS2 transistors with low-resistance Mo contacts [29c]. Density
profiles revealing sulfur incorporation and molybdenum reduction functional theory (DFT) simulations indicate Mo can form a high-
were studied. Tunable film thickness with centimeter-scale mono- quality contact interface with monolayer MoS2 with zero tunnel
layer growth was successfully achieved (Fig. 26) [28e]. In this study, barriers and zero Schottky barriers.
a higher temperature was required for the formations of 2-H MoS2 Recently, graphene has been used as an electrode for MoS2 tran-
crystal structure. sistor because it forms a subtle Schottky barrier. Graphene is well
known to possess high electrical conductivity, optical transparency,
5. Applications and flexibility, a flexible and transparent MoS2 transistor has also
been fabricated using graphene as source-drain electrodes [29d].
5.1. TMDCs transistors Lee et al. have compared graphene vs. metals source/drain con-
tact for MoS2 transistors and found that the mobility is higher for
TMDCs have unique features like stability in air, the absence of graphene contacted MoS2 as compare to metal contacted [29e-f].
a short-channel effect, no dangling bonds, and high mobility com- In another study, Roy et al. found that 2D materials for all com-
parable to the Si which makes them attractive as an ideal channel ponents, including MoS2 as the channel material, hexagonal BN as
material for FETs [12a,24,25]. The earliest application of TMDCs for top gated dielectric and graphene as electrodes works well [30b].
FETs was reported in 2004, where WSe2 crystals showed mobility In a similar study, Singh et al. demonstrated low-voltage, high-
comparable to the best single-crystal Si FETs (up to 500 cm2 V−1 s−1 performance field-effect transistors fabricated using single-layer
for p-type conductivity at room temperature), ambipolar behavior graphene with multilayered MoS2 as a channel (Fig. 28) [30c]. The
with a 104 on/off ratio at a temperature around 60 K [29a]. Kis and two terminal mobility of graphene contacted to monolayer MoS2
A.K. Singh et al. / Applied Materials Today 13 (2018) 242–270 261

Fig. 26. Multistep sulfurization of (a) Mo 3d XPS normalized by the maximum signal for as-deposited ALD and PEALD films following single step processing at 300 ◦ C, 2
step processing first at 300 ◦ C and then at 600/700 ◦ C, and 3 step processing at 300, 600, then 900 ◦ C. Dashed lines represent fitted contributions of S 2s and Mo4+ 3d5/2 , and
the solid line represents total contribution. (b) The Atomic percentage of sulfur based on the Mo 3d and S 2s peaks showing partial incorporation of sulfur during a 300 ◦ C
single step anneal and final stoichiometric incorporation via a three-step anneal. (c) S 2p XPS spectra with broad peaks at 300 ◦ C indicating disorder and poor uniformity in a
chemical environment. By 600 ◦ C in a two-step anneal, peaks are sharp (binding energies of 162.2 eV for S 2p3/2 and 163.5 for eV 2p1/2 respectively). Dashed lines indicate
fits for the 2p3/2 and 2p1/2. Solid lines represent the sum of the fits. As the sulfur content increases at higher temperature, Raman spectra (d) show the emergence of the A1g
and E2g peaks characteristic of the 2H structure of multilayer MoS2 . (e) Optical image of 4.5 nm thick MoS2 following transfer to a larger 300 nm SiO2 on Si substrate showing
centimeter scale uniformity achieved with ALD and the post sulfurization process. (Reproduced with permission from ref. 28(e). Chem. Mater., 2017, 29, 2024–2032.)

at room temperature using 15 nm Al2 O3 as the top-gate dielec- 10 ␮A/cm2 field emission current density for MoS2 layers was
tric layer was found to be 131.2 cm2 V−1 s−1 . The reported values reported to be 3.5 V/␮m (Fig. 29). The turn on values found to
are higher than the previously reported metal/graphene-contacted be notably lower than graphene. The emission current-time plots
MoS2 [30c]. Furthermore, Singh et al. tuned the threshold voltage, show excellent stability over the period of 3 h. Due to the weak turn-
charge carrier concentration and mobility of the single, bi-and mul- on field and planar (sheet-like) structure, MoS2 may have potential
tilayer transistors of MoS2 with the combination of gases under application in futuristic vacuum microelectronics and flat panel
ultraviolet light exposure [30d] and by implementing gold (Au) display [7d].
metal adsorbate [30c].
5.3. Photovoltaics and photodetection
5.2. Field emission
Photovoltaics and photodetection are another interesting filed
Field emission studies of layered MoS2 sheets at the base pres- where TMDCs thin films can be employed. A Schottky-barrier solar
sure of ∼1 × 10−8 mbar were investigated. A turn-on field to draw cell fabricated using few-layer, CVD-grown MoS2 film serving as
262 A.K. Singh et al. / Applied Materials Today 13 (2018) 242–270

cell is shown in the insets of Fig. 32(a) and (b), respectively [30f].
The study concluded that higher the MoS2 thickness resulting
higher the JSC values for a solar cell that is due to efficient photo-
absorption in the thicker MoS2 system which leads to a noteworthy
alteration in the overall device performance.
The reported results on solar cell performance would promote
continued efforts toward developing highly efficient Schottky-
barrier solar cells captivating advantage of the unique interfacial
properties of layered semiconductor nanostructures [30f].

5.4. Gas sensing

Recently, Dravid’s group reported on the synthesis of large-area


mono- to hex-layers MoS2 sheets and identified using an optical
microscope with the help of distinctly visible color contrast and
further confirmed by AFM and Raman spectroscopy [4e]. The var-
ious MoS2 sheets in the form of transistors were assessed for gas
sensing performance with exposure to NO2 , NH3 , and humidity, at
room temperature [4e]. Upon exposure of NO2 and NH3 gases, the
MoS2 transistors showed a decrease and increased in conductance,
respectively with a substantial shift in the threshold gate voltage.
Single-layer MoS2 shows quick response but the low signal to noise
ratio. The bi-to hex-layer MoS2 shows sound sensitivity as well as
proper intervention and recovery. These results indicate that 2-
to 6-layer MoS2 transistor-based devices exhibit excellent sensing
performance at room temperature, compared to the existing solid-
Fig. 27. (a) A cross-sectional view of the structure of a monolayer MoS2 FET. (b)
Transfer characteristic at room temperature for the FET with 10 mV applied bias state sensors, with better sensitivity, quick response, and ability to
voltage Vds. Backgate voltage Vbg is being implemented to the substrate, and the sense lower concentrations. Fig. 33 shows the comparative study
top gate is disconnected. The inset shows the Ids –Vds curve acquired for Vbg values of sensing behavior with and without applying back gate voltage
of 0, 1 and 5 V. (Reproduced with permission from ref. 14(d). Nature Nanotechnology, (+15 V) for 2-layer MoS2 (a) NH3 , (b) NO2 and 5-layer MoS2 (c) NH3
2011, 6, 147–150.)
and (d) NO2 [4e].

the critical photo-active layer is recently reported [30f]. Fig. 30 5.5. Hydrogen production
shows CVD growth setups to synthesize MoS2 on SiO2 /Si substrates.
Fig. 30(b–e) realizes the steps for transferring MoS2 films on the An ever increasing demand for energy in our society leads to the
desired substrates [30f]. Fig. 31 provides information on the (a) exploitation of non-renewable energy sources including fossil fuel
Schematic cross-sectional view of the Schottky barrier solar cell paved the way for several environmental crises [31a]. As a conse-
on an ITO substrate with Au contact. (b) Energy band diagram of quence of the intensive use of fossil fuels, global warming is one of
the solar cell with the formation of a Schottky-barrier between the the principal threats due to the accumulation of greenhouse gases.
MoS2 nanomembrane with Au metal contact. Limited availability of rapidly exhausting fossil fuels has compelled
The performance of solar cell depending on MoS2 thickness was researchers to accelerate the search for environment-friendly,
investigated and presented in Fig. 31. Fig. 32(a) and (b) shows dark renewable and sustainable alternative energy resources [31b-c].
and illuminated J–V characteristics of the as-fabricated Schottky- To lessen the dependence on fossil fuels, hydrogen is considered as
barrier solar cells with MoS2 thicknesses of 110 nm and 220 nm, a promising cleaner-energy alternative in comparison to carbon-
respectively. The output power as a function of voltage in a solar based fossil fuels as the combustion of hydrogen produces water

Fig. 28. (a) A cross-sectional view of the top-gated ML MoS2 , (b) Plot of IDS -Vtg of the ML MoS2 transistor at VDS value of 0.01 V. (Reproduced with permission from ref. 30(c).
ACS Appl. Mater. Interfaces, 2016, 8, 34699–34705.)
A.K. Singh et al. / Applied Materials Today 13 (2018) 242–270 263

Fig. 29. Field emission characteristics of a multilayer MoS2 system, (a) represent the applied electrical field behavior (J) as a function of current density (E). And (b) F-N
plot showing non-linear behavior indicating emission current from the semiconducting emitter, the inset shows field emission pattern of a few layered MoS2 recorded at a
current density of 50 ␮A/cm2 . (Reproduced with permission from ref. 7(d) Small, 2013, 9, 2730–2734.)

Fig. 30. (a) Schematic of the CVD growth setup of a MoS2 nanomembrane on Si-SiO2 substrate. (b) Major process steps involved in the synthesis of the MoS2 nanomembrane,
layer transfer, and subsequent fabrication of a Schottky-barrier solar cell. The KOH wet etching process was used to detach the MoS2 nanomembrane from SiO2 –Si substrate.
(c and d) Partially floating MoS2 nanomembrane in KOH solution. (e) Free-floating MoS2 nanomembrane in KOH to be transferred onto an ITO-coated glass substrate for solar
cell fabrication. (Reproduced with permission from ref. 30(f), Nanoscale 2012, 4, 7399–7405.)
264 A.K. Singh et al. / Applied Materials Today 13 (2018) 242–270

Fig. 32. Measured dark and illuminated J–V characteristics of the Schottky-barrier
Fig. 31. (a) Schematic cross-sectional view of the demonstrated Schottky barrier
solar cells with a stack of MoS2 nanomembrane having thicknesses of (a) 110 nm
solar cell structure showing a stack of MoS2 nanomembrane on an ITO substrate
and (b) 220 nm. Inset of each figure shows the power output as a function of voltage
with Au contact. (b) Energy band diagram of the solar cell with the formation of a
for the fabricated solar cell. (Reproduced with permission from ref. 30(f) Nanoscale
Schottky-barrier between the MoS2 nanomembrane and Au metal contact. (Repro-
2012, 4, 7399–7405.)
duced with permission from ref. 30(f) Nanoscale 2012, 4, 7399–7405.)

required, which corresponds to a thermodynamic electrochemical


as the only by-product without emission of any toxic or green-
potential of 1.23 V [8i]. Therefore, an external potential of ∼1.23 V
house gases. Hydrogen production using environment-friendly
is sufficient (theoretically) for water splitting in an electrochemical
photo/electro/chemical methods is critically important and signifi-
cell assembly. However, the substantial kinetic barriers originat-
cantly promising approach for the utilization of affordable, clean
ing from the high activation energies required for the formation
energy [31d-f]. Water splitting either by (1) electrocatalytic or
of reaction intermediates on the electrode surface which leads to
(2) photo-electrocatalytic route requires the hydrogen evolution
overpotentials. Thus, for practical realization, a potential more than
reaction (HER) to occur readily. Since the emergence of photo-
1.23 V is needed for water splitting [8i].
catalytic water splitting in 1972 [31g], HER received significant
The HER in an acidic environment typically involves three steps
research attention. The production of hydrogen involves the uti-
[32a]: (i) The first step is called the Volmer or discharge reaction
lization of precious noble metals (e.g., Pt, Rh, Pd, etc.) and their
with a Tafel slope of 120 mV dec−1 and can be presented as:
alloys [8i]. However, high cost and inadequacy, significantly hin-
dering their industrial application. The scalable and sustainable H+ (aq) + e− H∗ads
hydrogen production demands efficient and robust earth-abundant
During this reaction between an electron and proton transfer,
electrocatalysts for the HER beyond precious-metal. Therefore,
on the electrode surface generation and adsorbtion of hydrogen
development of low-cost and earth-abundant catalysts as a com-
atom (H∗ads ) take place. The intermediate reaction could proceed
petitive alternative to Pt-group metals is highly desirable.
two ways: either via the (ii) Tafel reaction or (iii) the Heyrovsky
reaction.
6. Hydrogen evolution mechanism In the Tafel reaction, two absorbed hydrogen atoms combine to
generate H2 gas with a Tafel slope of 30 mV dec−1 :
The water-splitting reaction:
2H∗ads H2 (gas)
H2 O H2 + 1/2 O2
Whereas Heyrovsky reaction takes place when another elec-
Can be divided into two half-reactions: the water oxidation reac- tron reacts with an adsorbed hydrogen atom and a proton from the
tion (or oxygen evolution reaction [OER]) and the water reduction solution to yield H2 with a Tafel slope of about 40 mV dec−1 [32b,c].
reaction (or HER). For this reaction to take place and to achieve
the electrochemical water splitting, an energy G = 237.1 kJ/mol is H∗ads + H+ (aq) + e− H2 (gas)
A.K. Singh et al. / Applied Materials Today 13 (2018) 242–270 265

Fig. 33. Comparative sensing behaviors with and without applying back gate voltage (+15 V) for 2-layer MoS2 (a) NH3 , (b) NO2 and 5-layer MoS2 (c) NH3 and (d) NO2 .
(Reproduced with permission from ref. 4(e) ACS Nano, 2013, 7, 4879–4891.)

It is quite evident from the reaction mechanisms that an including (a) increasing active sites, (b) tuning the phase and
adsorbed hydrogen (H∗ads ) as an intermediate is of vital importance. electronic structure, (c) coupling with conductive scaffolds, (d)
Since H∗ads is always generated in the first Volmer reaction con- amorphous molybdenum sulfide, and (e) (amorphous) ternary
secutively, takes part in each electrochemical reaction step in the compounds demonstrating meritoriously improved catalytic activ-
course of HER. Furthermore, it has been well established that the ity of MoS2 [33f-i]. Inspired from the fact that the higher catalytic
H∗ads energies for HER are moderated by optimum catalysts fea- activity can be achieved along the edges of MoS2 , various nanos-
ture [32d]. The theoretical calculations providing insights for the tructures of MoS2 including nanosheets, nanoflakes, nanowires,
catalytic activity with the help of density functional theory (DFT) nanoparticles and mesoporous structures, have been fabricated
while computing the Gibbs free energy for hydrogen adsorption to optimize the exposed active sites [33f-i]. It has been well
( GH *) in the Volmer reaction has been proposed as an effective established that the single-layered MoS2 particularly with Mo-
tool for HER catalytic activity [32e-g]. Pt-group metals with almost terminated edges, contribute predominantly to the HER catalytic
zero GH * and the highest exchange current density are the best activities [36a,b]. Theoretical calculations based on DFT simu-
suitable materials for the HER, but due to several restrictions as lation reveals that most of the HER activity is ascribed to the
mentioned above, their practical realization is hindered. Therefore, Mo-terminated edge MoS2 owning lower hydrogen binding energy
a non-Pt earth-abundant HER catalysts with nearly zero GH * is than the basal plane [32f,36c-f]. Some recent study based on DFT
highly desired. Significant research efforts have been devoted in the calculations claimed that the sulfur vacancy is the new catalytic site
exploration of efficient, inexpensive, earth-abundant, and nontoxic for HER in 2H-MoS2 beyond the edges [36g].
catalysts for HER, e.g., metal alloys, chalcogenides, nitrides, phos- Vertically aligned layers of MoS2 and MoSe2 thin films provid-
phides, borides, carbides and their composites etc. [33,34]. Among ing maximum exposure of the edges on the film surface have been
several, two-dimensional (2D) layered transition-metal dichalco- reported by Kong et al. [37a] (Fig. 34(a-b)). In a different approach,
genides (TMDCs) have received great attention because of their MoS2 nanoflowers grown on graphite substrates by chemical vapor
unusual physical/chemical properties originating from their unique deposition (CVD) method successfully exposed the abundant edges
crystal structures, with great promise for hydrogen production as [37b]. Single-crystal MoS2 nanobelts exhibiting high electrocat-
HER catalysts [35a-c]. The last decade has witnessed the develop- alytic hydrogen evolution efficiency have also been reported [37c].
ment of Molybdenum disulfide (MoS2 ) conceivably the first earth Kibsgaard and co-worker adopted an elaborate strategy altering the
abundant compound to be studied as a promising candidate of MoS2 surface to improve catalytic performance (Fig. 34(c-f)) [36c].
noble-metal-free HER catalyst [33d,f-i,35c-e]. MoS2 with low cost, In another experiment, ultrathin MoS2 nanosheets with a large
earth abundance, and excellent stability is suggested as a promis- number of defects were synthesized delivering improved catalytic
ing HER catalyst because of low GH * close to that of noble metals performance [36d].
[33f-i]. It is well documented that the bulk MoS2 is not HER active
catalyst [35f]. However; theoretical calculations predicted that the 7. Composites with conductive hosts
nanostructures of MoS2 could serve as a promising HER catalyst
[32f]. A composite of MoS2 with conductive frameworks is an
Recently, several crucial aspects have been developed to sig- effective method to improve the overall HER catalytic per-
nificantly enhance HER catalytic activity of nanostructured MoS2 formance [33f-i]. Different conductive carbon nanostructures
266 A.K. Singh et al. / Applied Materials Today 13 (2018) 242–270

Fig. 34. (a) TEM image of a MoS2 film fabricated by rapid sulfurization process. (b) Idealized structure of edge-terminated molybdenum chalcogenide films with maximally
exposing the edges of the layers. Reprinted with permission from Kong et al. 37(a) Copyright 2013 American Chemical Society. (c) Schematic for the synthesis of double-gyroid
mesoporous MoS2 engineered to preferentially expose edge sites for enhanced HER activity. (d) TEM image of thus synthesized MoS2 film (e) Tafel slope of double-gyroid
MoS2 v/s core-shell MoO3 -MoS2 nanowires, showing the 50 mV per decade slope once corrected for the resistance of the MoS2 . (f) Ratios of surface area, density of active sites
per surface area and total HER activity of the various double-gyroid MoS2 films v/s the nanowires. (Reprinted by permission from Macmillan Publishers Ltd: Nature Materials
(Kibsgaard et al. 36(c)), copyright 2012.)

Fig. 35. (a) Scheme for the solvothermal synthesis, (b) SEM and TEM (inset) images of MoS2 -graphene nanocomposite. (c) Polarization curves, (d) Tafel plots recorded on
glassy carbon electrodes with a catalyst loading of 0.28 mg/cm2 and (e) Durability test for the MoS2 /RGO hybrid catalyst. Reprinted with permission from Li et al. 36(a) Copyright
2011 American Chemical Society.

including graphene, graphene oxide, carbon nanotubes, and of ∼41 mV dec−1 . The superior performance was observed due
carbon nanofibers, etc. have been most common hosts to fabri- to the excellent electrical coupling of the RGO network with
cate integrated MoS2 nanostructures [33f-i]. Dai and coworkers MoS2 . Moreover, MoS2 -based hybrid nanocomposites were pre-
reported a selective solvothermal process to synthesize MoS2 pared to utilize graphene foam, carbon nanotubes, and carbon
nanoparticles on reduced graphene oxide (RGO) sheets sus- nanofiber as conducting hosts exhibiting excellent catalytic activ-
pended in solution (Fig. 35) [36a]. Thus synthesized, RGO@MoS2 ity [32b,37d]. In a similar experiment, the graphene-deposited
composite shows excellent electrocatalytic activity at a low 3D-Ni foam was used to fabricate MoSx catalysts for HER appli-
catalytic overpotential of ∼0.1 V with a smaller Tafel slope cations.
A.K. Singh et al. / Applied Materials Today 13 (2018) 242–270 267

8. Defective MoS2 nanosheets and amorphous MoS2 by a photovoltaics-electrolyzer system or an integrated PEC
water-splitting system [31d–f]. The integrated PEC systems for
Defect engineering and growth of amorphous molybdenum water splitting at the semiconductor interface referred to be more
sulfide (MoSx ) are another efficient way to increase the HER economical promising better efficiency. A typical PEC assembly is
catalytic activity [36d,h]. The defective MoS2 nanosheets with composed of two essential components: (1) a light absorber (usu-
rich, active sites reported to exhibit improved catalytic activity. ally a semiconductor) to produce electron-hole pairs upon light
On the other hand, amorphous molybdenum sulfide (MoSx ) pre- illumination and (2) an electrocatalyst responsible for the charge
serve short-range ordering with the substantial structural disorder; transfer and to reduce the overpotential for hydrogen production.
the high surface area and structural defects lead to possess- In a PEC assembly, water splitting could be achieved with either a
ing more catalytic active sites for HER [33e]. The defect-rich single semiconductor (bandgap higher than ∼1.7 eV) or a combi-
MoS2 exhibited a smaller Tafel slope, lower onset overpoten- nation of two or more semiconductors in a tandem PEC cell [8i].
tial (∼120 mV) and larger cathodic current density as compared A typical tandem PEC cell consists of a hydrogen-evolving photo-
to defect-free MoS2 [36d,38a]. The plasma treatment was sug- cathode and an integrated oxygen-evolving photoanode [31d-f]. A
gested as an efficient way to fabricate defect-rich MoS2 [38b,c] tandem PEC cell comprising more than one semiconducting mate-
for HER application. The amorphous molybdenum sulfide (MoSx ) rials with lower bandgaps enables access to a more substantial
was successfully synthesized via electrodeposition or wet chemi- part of the solar spectrum as compared to a single semiconductor,
cal reactions demonstrating excellent HER activity [38d,e]. Merki thus greatly preferred. The application of TMDCs for photoelec-
and co-worker reported the effect of transition-metal ions (Mn, trocatalytic HER has long been investigated [31g,8i,32a]. The
Fe, Ni, Co, Cu, and Zn) doping on the HER activity of MoSx semiconducting TMDCs possessing a direct band gap between
[39a]. It is hard to characterize the atomic structures precisely, 1.4–2.3 eV and an indirect band gap of 1.0–1.5 eV will absorb the
categorizing the active catalytic sites, to understand the over- incident photon with energies equal or above direct/indirect band
all catalytic mechanism(s) for amorphous MoSx catalysts. Even gap, and show large absorption coefficients (≈106 cm−1 ) [41]. The
after several attempts, the reaction mechanism of HER activity for heterostructures of TMDCs and metal are reported to effectively
amorphous molybdenum sulfide (MoSx ) is not fully understood enhance the HER activity of a PEC cell assembly due to an increase in
[38d,39b–e]. the local density of states near the TMDCs/metal interface [41a–d].
Moreover, 1T-MoS2 exhibit significantly enhanced HER catalytic
activity as compared to 2H-MoS2 [37b]. An integrated PEC cell
9. Role of the phase, edge, and vacancies
assembly of planar p-type Si and 1T-MoS2 (1T-MoS2 @Si) is reported
serving as an efficient and robust system for PEC hydrogen produc-
Recently, more insights have been provided on the HER catalytic
tion [41b]. Furthermore, it was confirmed that the charge-transfer
activity using mesoporous (holey) 1T-phase MoS2 nanosheets (P-
resistances at the semiconductor/catalyst and catalyst/electrolyte
1T-MoS2 ) synthesized through a liquid ammonia-assisted lithiation
interfaces of 1T-MoS2 /Si is superior to 2H-MoS2 . In another exper-
route [40a]. The report successfully investigated and compare the
iment, Lewis and co-worker reported enhanced PEC performance
contributions of crystal structure (phase), edges, and sulfur vacan-
of Pt-decorated p-type WSe2 photocathodes [42a]. A more detailed
cies (S vacancies) with mesoporous 2H-phase MoS2 , mesoporous
and systematic study performed by Chang et al. [42b] examined the
2H-phase MoS2 after sulfur compensation, 1T-phase MoS2 , and 2H-
relationship between a number of MoS2 layer and photocatalytic
phase MoS2 [40a]. Furthermore, it was revealed that: (1) the 1T
hydrogen generation. As it is well known that the layer number
phase MoS2 exhibits superior HER catalytic activity over the 2H
of MoS2 plays an important role of HER activity. In the experi-
phase; (2) an increase in the density of edges, can further improve
ment designed by Chang et al. [42b], MoS2 layers varying from 1
the catalytic performance; (3) P-1T-MoS2 deliver better perfor-
to 112 were synthesized loaded with CdS (MoS2 @CdS). The cor-
mance as compared to 1T-MoS2 ; and (4) the contribution of S
responding hydrogen production activities were explored in lactic
vacancies to HER catalysis should not be neglected. As a result from
acid and Na2 S-Na2 SO3 solutions. From the study, it is observed that
S vacancies and edges, high intrinsic HER activity can be obtained
the highest H2 production rate was achieved with MoS2 @CdS has
on porous 1T-MoS2 nanosheets, with a Tafel slope ∼43 mV dec−1 ,
a single-layer assembly confirming that the photocatalytic activ-
which is among the highest performances reported for phase-pure
ity is critically dependent on the number of layer participating
MoS2 . In a recent report by Zhang et al. the HER activities at the
(increasing with decreasing MoS2 layer number) [42b]. Excitingly
edge and basal-plane sites of single-layer MoS2 synthesized by
it was noted that the H2 production rate using MoS2@CdS (single
CVD method were carried out using a local probe method enabled
layer assembly) in lactic acid solution (2.59 mmol h−1 ) is higher
by selected-area lithography. The e-beam lithography was utilized
than that in the Na2 S-Na2 SO3 solution (2.01 mmol h−1 ) which
to open reaction windows at sites of interest on poly (methyl
is much higher as compared to Pt/CdS in lactic acid solution
methacrylate) (PMMA) covered single layered MoS2 triangles. The
(0.44 mmol h−1 ) [42c]. In a very recent study, a ternary hybrid
study revealed the improvements in HER activities (i.e., decreased
Cu2 ZnSnS4 (CZTS)/MoS2 -reduced graphene oxide (RGO) is reported
overpotential and lower the Tafel slope) are resultant from the
as the promising alternative of noble metals for H2 generation.
active sites of Mo-terminated edge and 2H to a 1T phase tran-
The heterostructure exhibit exceptionally higher (∼320%) HER
sition. Additionally, it was concluded that the active sites on the
activity than bare CZTS as well as Au or Pt decorated CZTS too
Mo-terminated edge of both 2H- and 1T -MoS2 were critical with
[42d].
optimum HER performance, and the basal-plane activity of 1T -
MoS2 was favorable for the HER in comparison with 2H counterpart
[40a].
11. Prospects and direction

10. The photoelectrocatalytic (PEC) HER This brief review is an attempt to highlight the various prop-
erties and theoretical aspects of MoS2 and its use in the field of
Photoelectrochemical (PEC) water splitting is considered as 2D nanoelectronic devices. Although limited progress has been
another one of the most auspicious methodologies besides the made in discovering new QSH insulators based on MoS2 dur-
electrocatalysis to produce hydrogen fuel [31g,8i,32a]. In the PEC ing the recent years, the implementation of robust tools for
HER mechanism, the hydrogen production can be achieved either assessing topological characteristics of band structures (e.g., Z2
268 A.K. Singh et al. / Applied Materials Today 13 (2018) 242–270

invariant) will help to accelerate the discovery process for vari- References
ous technological applications. The primary challenge in the field
of topological materials is to increase the gap to realize high- [1] (a) K.S. Novoselov, A. Mishchenko, A. Carvalho, A.H.C. Neto, Science 353
(2016) 9439;
temperature topological state, which can be met by designing (b) A.K. Geim, K.S. Novoselov, Nat. Mater. 6 (2007) 183–191;
new allotropic forms of MoS2 . Also, the computer mimicking (c) A.K. Geim, Science 324 (2009) 1530–1534;
is increasingly used to explore the physics of novel TIs, where (d) K.S. Novoselov, A.K. Geim, S.V. Morozov, D. Jiang, Y. Zhang, S.V. Dubonos,
I.V. Grigorieva, A.A. Firsov, Science 306 (2004) 666;
new candidates of topological materials are often predicted and (e) D.J. Late, A. Ghosh, K.S. Subrahmanyam, L.S. Panchakarla, S.B. Krupanidhi,
designed for experimental synthesis and characterization. DFT- C.N.R. Rao, Solid State Commun. 150 (2010) 734–738;
based electronic structure calculations are expected to provide (f) B.H. Nguyen, V.H. Nguyen, Adv. Nat. Sci.: Nanosci. Nanotechnol. 7 (2016)
023002;
mechanistic insights of the novel TIs based on MoS2 and future (g) G. Jo, M. Choe, S. Lee, W. Park, Y.H. Kahng, T. Lee, Nanotechnology 23
promises for the applications in quantum computing and spin- (2012) 112001.
tronics. It seems that MoS2 is just star material for investigation [2] (a) P. Kumar, A.K. Singh, S. Hussain, K.N. Hui, K.S. Hui, J. Eom, J. Jung, J. Singh,
Rev. Adv. Sci. Eng. 2 (2013) 238–258;
of similar materials to be used in various device applications
(b) A.K. Singh, M. Ahmad, V.K. Singh, K. Shin, Y. Seo, J. Eom, ACS Appl. Mater.
shortly. Interfaces 5 (2013) 5276;
(c) A.K. Singh, M.W. Iqbal, V.K. Singh, M.Z. Iqbal, J.H. Lee, S.H. Chun, K. Shin, J.
Eom, Mater. Chem. 22 (2012) 15168;
12. Summary
(d) A.K. Singh, J. Eom, ACS Appl. Mater. Interfaces 6 (2014) 2493–2496;
(e) S. Pang, Y. Hernandez, X. Feng, K. Müllen, Adv. Mater. 23 (2011)
In brief, this review is an attempt to highlight the various prop- 2779–2795;
erties and theoretical aspects of MoS2 summering its potential (f) J. Plutnar, M. Pumera, Z. Sofer, J. Mater. Chem. C (2018),
doi:10.1039/C8TC00463C.
applications in many fields. The excellent electronic, spintronics, [3] (a) T.H. Han, Y. Lee, M. Choi, S.H. Woo, S.H. Bae, B.H. Hong, J.H. Ahn, T.W. Lee,
sensing and catalytic performance suggested that MoS2 is a pos- Nat. Photon. 6 (2012) 105–110;
sible substitute for the state of the art materials currently being (b) W. Yuan, A. Liu, L. Huang, C. Li, G. Shi, Adv. Mater. 25 (2013) 766–771;
(c) A.D. Smith, F. Niklaus, A. Paussa, S. Vaziri, A.C. Fischer, M. Sterner, F.
used in several industries. Even though MoS2 is an earth-abundant Forsberg, A. Delin, D. Esseni, P. Palestri, M. Östling, M.C. Lemme, Nano Lett. 13
material, reliable and high-quality production, from the fundamen- (2013) 3237–3242;
tal to scalable technological applications is extensively demanded. (d) A.K. Singh, S. Andleeb, J. Singh, H.T. Dung, Y. Seo, J. Eom, Adv. Func. Mater.
24 (2014) 7125–7132;
Challenges including large-scale production, cost, and stability of (e) B. Radisavljevic, M.B. Whitwick, A. Kis, ACS Nano 5 (2011) 9934–9938;
MoS2 are critically important. It is obvious to overcome the chal- (f) M. Begliarbekov, S. Strauf, C.P. Search, Nanotechnology 22 (2011) 165203;
lenges, reducing the cost will pave the way for mass production (g) D. Dutta, A. Hazra, S.K. Hazra, J. Das, S. Bhattacharyya, C.K. Sarkar, S. Basu,
Meas. Sci. Technol. 26 (2015) 115104;
while long-term stability guarantees industrial application. Limited (h) S.J. Jeong, H.W. Kim, J. Heo, M.H. Lee, H.J. Song, J. Ku, Y. Lee, Y. Cho, W. Jeon,
progress has been made in discovering new QSH insulators based H. Suh, S. Hwang, S. Park, 2D Mater. 3 (2016) 035027;
on MoS2 during the recent years, the implementation of robust (i) J.H. Garcia, Marc Vila, A.W. Cummings, S. Roche, Chem. Soc. Rev. (2018),
DOI: 10.1039/C7CS00864C.
tools for assessing topological characteristics of band structures
[4] (a) O.L. Sanchez, D. Lembke, M. Kayci, A. Radenovic, A. Kis, Nat. Nanotechnol. 8
(e.g., Z2 invariant) will help to accelerate the discovery process (2013) 497–501;
for various technological applications. The primary challenge in (b) W. Zhang, J.K. Huang, C.H. Chen, Y.H. Chang, Y. Cheng, L.J. Li, Adv. Mater. 25
the field of topological materials is to increase the gap to realize (2013) 3456–3461;
(c) S. Bertolazzi, D. Krasnozhon, A. Kis, ACS Nano. 7 (2013) 3246;
high-temperature topological state, which can be met by design- (d) H. Li, Z. Yin, Q. He, H. Li, X. Huang, G. Lu, D.W.H. Fam, A.I.Y. Tok, Q. Zhang,
ing new allotropic forms of MoS2 . Also, the computer mimicking H. Zhang, Small 8 (2012) 63–67;
is increasingly used to explore the physics of novel TIs, where (e) D.J. Late, Y.K. Huang, B. Liu, J. Acharya, S.N. Shirodkar, J. Luo, A. Yan, D.
Charles, U.V. Waghmare, V.P. Dravid, C.N.R. Rao, ACS Nano. 7 (2013)
new candidates of topological materials are often predicted and 4879–4891;
designed for experimental synthesis and characterization. DFT- (f) P.J. Ko, A. Abderrahmane, N.H. Kim, A. Sandhu, Semicond. Sci. Technol. 32
based electronic structure calculations are expected to provide (2017) 065015.
[5] (a) Y. Hang, B. Zheng, C.F. Zhu, X. Zhang, C.L. Tan, H. Li, B. Chen, J. Yang, J.Z.
mechanistic insights of the novel TIs based on MoS2 and future Chen, Y. Huang, L.H. Wang, H. Zhang, Adv. Mater. 27 (2014) 935–939;
promises for the applications in quantum computing and spintron- (b) K. Lee, R. Gatensby, N. McEvoy, T. Hallam, G.S. Duesberg, Adv. Mater. 25
ics. (2013) 6699;
(c) M. Acerce, D. Voiry, M. Chhowalla, Nat. Nanotech. 10 (2015) 313–318;
Although MoS2 -based catalysts have shown promising poten- (d) K. Leng, Z.X. Chen, X.X. Zhao, W. Tang, B.B. Tian, C.T. Nai, W. Zhou, ACS
tials as electrochemical HER catalyst, the development in Nano 10 (2016) 9208–9215;
photoelectrochemical hydrogen production is relatively slower. To (e) L. Cao, S. Yang, W. Gao, Z. Liu, Y. Gong, L. Ma, G. Shi, S. Lei, Y. Zhang, S.
Zhang, R. Vajtai, P.M. Ajayan, Small 9 (2013) 2905;
the direct realization of solar-to-fuel conversion using MoS2 -based
(f) M.W. Lin, L. Liu, Q. Lan, X. Tan, K.S. Dhindsa, P. Zeng, V.M. Naik, M.M.C.
catalysts, the more rigorous efforts are needed. Even after the sig- Cheng, Z. Zhou, J. Phys. D: Appl. Phys. 45 (2012) 345102.
nificant improvement in overall catalytic performance, the HER [6] (a) S. Zhao, T. Hotta, T. Koretsune, K. Watanabe, T. Taniguchi, K. Sugawara, T.
catalytic activity of MoS2 is still unable to beat Pt-based noble met- Takahashi, H. Shinohara, R. Kitaura, 2D Mater. 3 (2016) 025027;
(b) S. Xu, Z. Wu, H. Lu, Y. Han, G. Long, X. Chen, T. Han, W. Ye, Y. Wu, J. Lin, J.
als. The in-depth understanding of the catalytic mechanisms on the Shen, Y. Cai, Y. He, F. Zhang, R. Lortz, C. Cheng, N. Wang, 2D Mater. 3 (2016)
atomic level is not yet fully defined. It seems that MoS2 is just a star 021007;
material, it is strongly suggested that the theoretical and experi- (c) T. Paul, S. Ghatak, A. Ghosh, Nanotechnology 27 (2016) 125706;
(d) J. Ahn, P.J. Jeon, S.R.A. Raza, A. Pezeshki, S.W. Min, D.K. Hwang, S. Im, 2D
mental works should be integrated to enable more efficient rational Mater. 3 (2016) 045011;
design for various applications. (e) F. Liu, J. Wang, H. Guo, Nanotechnology 26 (2015) 175201.
[7] (a) D.J. Late, B. Liu, H.S.S.R. Matte, C.N.R. Rao, V.P. Dravid, Adv. Funct. Mater. 22
(2012) 1894–1905;
Acknowledgments (b) D.J. Late, B. Liu, J.J. Luo, A.M. Yan, H.S.S.R. Matte, M. Grayson, C.N.R. Rao,
V.P. Dravid, Adv. Mater. 24 (2012) 3549–3554;
This work was supported by the Department of Science & (c) H.S. Matte, A. Gomathi, A.K. Manna, D.J. Late, R. Datta, S.K. Pati, C.N.R. Rao,
Angew. Chem. Int. 49 (2010) 4059–4062;
Technology (DST), Government of India, and partially by National (d) R.V. Kashid, D.J. Late, S.S. Chou, Y. Huang, M. De, D. Joag, M.A. More, V.P.
Chemical Laboratory Project MLP-028626. Jai Singh would like to Dravid, Small 9 (2013) 2730–2734;
acknowledge UGC-India and DST for providing project under UGC (e) S. Bae, H. Kim, Y. Lee, X. Xu, J. Park, Y. Zheng, J. Balakrishnan, T. Lei, R.H.
Kim, Y. Song, Y.J. Kim, K.S. Kim, B. Özyilmaz, J.H. Ahn, B.H. Hong, S. Iijima, Nat.
Start-up Grant and DST Fast track. Nanotechnol. 5 (2010) 574–578.
A.K. Singh et al. / Applied Materials Today 13 (2018) 242–270 269

[8] (a) D.J. Late, B. Liu, H.S.S.R. Matte, V.P. Dravid, C.N.R. Rao, ACS Nano 6 (2012) [17] (a) K.F. Mak, C. Lee, J. Hone, J. Shan, T.F. Heinz, Phys. Rev. Lett. 105 (2010)
5635–5641; 136805;
(b) F. Wang, Z. Wang, Q. Wang, F. Wang, L. Yin, K. Xu, Y. Huang, J. He, (b) A. Kuc, N. Zibouche, T. Heine, Phys. Rev. B 83 (2011) 245213;
Nanotechnology 26 (2015) 292001; (c) A. Kumar, P.K. Ahluwalia, Eur. Phys. J. B 85 (2012) 186;
(c) Y.C. Lin, D.O. Dumcencon, Y.S. Huang, K. Suenaga, Nat. Nanotechnol. 9 (d) A. Kumar, P.K. Ahluwalia, Mater. Chem. Phys. 135 (2012) 755;
(2014) 391–396; (e) A. Splendiani, L. Sun, Y. Zhang, T. Li, J. Kim, C.Y. Chim, G. Galli, F. Wang,
(d) R. Kappera, D. Voiry, S.E. Yalcin, B. Branch, G. Gupta, A.D. Mohite, M. Nano Lett. 10 (2010) 1271;
Chhowalla, Nat. Mater. 13 (2014) 1128–1134; (f) G.R. Bhimanapati, Z. Lin, V. Meunier, Y. Jung, J. Cha, S. Das, D. Xiao, Y. Son,
(e) N.A.H. Castro, F. Guinea, N.M.R. Peres, K.S. Novoselov, A.K. Geim, Rev. Mod. M.S. Strano, X. Valentino, R. Cooper, O.L. Liang, S.G. Louie, E. Ringe, W. Zhou,
Phys. 81 (2009) 109–162; S.S. Kim, R.R. Naik, B.G. Sumpter, H. Terrones, F. Xia, Y. Wang, J. Zhu, D.
(f) D.S. Schulman, A. Sebastian, D. Buzzell, Y.T. Huang, A.J. Arnold, S. Das, ACS Akinwande, N. Alem, J.A. Schuller, R.E. Schaak, M. Terrones, J.A. Robinson, ACS
Appl. Mater. & Interfaces 9 (2017) 44617–44624; Nano 9 (2015) 11509.
(g) Z. He, W. Que, Appl. Mater. Today 3 (2016) 23–56; [18] (a) J.B. Sun, J.B. Hannon, R.M. Tromp, P. Johari, A.A. Bol, V.B. Shenoy, K. Pohl,
(h) E. Benavente, M. Santa Ana, F. Mendizábal, G. González, Coord. Chem. Rev. ACS Nano 4 (2010) 7073;
224 (2002) 87–109; (b) W.Y. Liang, S.L. Cundi, Phis. Mag. 19 (1969) 1031;
(i) Q. Ding, B. Song, P. Xu, S. Jin, Chem 1 (2016) 699–726; (c) J.M. Soler, E. Artacho, J.D. Gale, A. Garcia, J. Junquera, P. Ordejon, D.
(j) R.J. Toh, Z. Sofer, J. Luxa, D. Sedmidubský, M. Pumera, Chem. Commun. 53 Sanchez-Portal, J. Phys.: Condens. Matter 14 (2002) 2745;
(2017) 3054–3057. (d) A. Kumar, P.K. Ahluwalia, Physica B 407 (2012) 4627;
[9] (a) Q.H. Wang, K. Kalantar-Zadeh, A. Kis, J.N. Coleman, M.S. Strano, Nat. Nano. (e) P. Johari, V.B. Shenoy, ACS Nano 5 (2011) 5903;
7 (2012) 699–712; (f) J.N. Coleman, M. Lotya, A. O’Neill, S.D. Bergin, P.J. King, U. Khan, K. Young, A.
(b) K.F. Mak, C. Lee, J. Hone, J. Shan, T.F. Heinz, Phys. Rev. Lett. 105 (2010) Gaucher, S. De, R.J. Smith, I.V. Shvets, S.K. Arora, G. Stanton, H.Y. Kim, K. Lee,
136805; G.T. Kim, G.S. Duesberg, T. Hallam, J.J. Boland, J.J. Wang, J.F. Donegan, J.C.
(c) A. Kuc, N. Zibouche, T. Heine, Phys. Rev. B 83 (2011) 245213; Grunlan, G. Moriarty, A. Shmeliov, R.J. Nicholls, J.M. Perkins, E.M. Grieveson, K.
(d) K. Kobayashi, J. Yamauchi, Phys. Rev. B 51 (1995) 17085–17095; Theuwissen, D.W. McComb, P.D. Nellist, V. Nicolosi, Science 331 (2011) 568;
(e) L. Liu, S.B. Kumar, Y. Ouyang, J. Guo, IEEE Trans. Electron Devices 58 (2011) (g) D. Penn, Phys. Rev. B 128 (1962) 2093.
3042–3047. [19] (a) K. Klitzing, G. Dorda, M. Pepper, Phys. Rev. Lett. 45 (1980) 494;
[10] (a) Y. Ding, Y. Wang, J. Ni, L. Shi, S. Shi, W. Tang, Physica B 406 (2011) (b) D.J. Thouless, M. Kohmoto, M.P. Nightingale, M. den-Nijs, Phys. Rev. Lett.
2254–2260; 49 (1982) 405–408;
(b) S. Lebègue, O. Eriksson, Phys. Rev. B 79 (2009) 115409; (c) F.D.M. Haldane, Phys. Rev. Lett. 61 (1988) 2015–2018;
(c) A. Splendiani, L. Sun, Y. Zhang, T. Li, J. Kim, C.Y. Chim, G. Galli, F. Wang, (d) C.L. Kane, E.J. Mele, Phys. Rev. Lett. 95 (2005) 226801;
Nano Lett. 10 (2010) 1271–1275; (e) C.L. Kane, E.J. Mele, Phys. Rev. Lett. 95 (2005) 146802;
(d) R.F. Frindt, A.D. Yoffe, Proc. R. Soc. Lond. A 273 (1963) 69–83; (f) L. Fu, C.L. Kane, E.J. Mele, Phys. Rev. Lett. 98 (2007) 106803;
(e) D.C. Elias, R.V. Gorbachev, A.S. Mayorov, S.V. Morozov, A.A. Zhukov, P. (g) M.Z. Hasan, C.L. Kane, Rev. Modern Phys. 82 (2010) 3045–3067.
Blake, L.A. Ponomarenko, I.V. Grigorieva, K.S. Novoselov, F. Guinea, A.K. Geim, [20] (a) B.A. Bernevig, T.A. Hughes, S.C. Zhang, Science 314 (2006) 1757–1761;
Nat. Phys. 7 (2011) 701–704; (b) D. Hsieh, D. Qian, L. Wray, Y. Xia, Y.S. Hor, R.J. Cava, M.Z. Hasan, Nature 452
(f) R.F. Frindt, Phys. Rev. 140 (1965) A536–A539; (2008) 970;
(g) R.F. Frindt, J. Appl. Phys. 37 (1966) 1928–1929; (c) L. Fu, C.L. Kane, Phys. Rev. B 76 (2007) 045302;
(h) F. Wang, Z. Wang, Q. Wang, F. Wang, L. Yin, K. Xu, Y. Huang, J. He, (d) L. Kou, Y. Ma, Z. Sun, T. Heine, T. Chen, J. Phys. Chem. Lett. 8 (2017)
Nanotechnology 26 (2015) 292001; 1905–1919.
(i) F. Bonaccorso, A. Lombardo, T. Hasan, Z. Sun, L. Colombo, A.C. Ferrari, [21] (a) G. Eda, T. Fujita, H. Yamaguchi, D. Voiry, M. Chen, M. Chhowalla, ACS Nano
Mater. Today 15 (2012) 564–589. 6 (2012) 7311;
[11] (a) A.S. Mayorov, R.V. Gorbachev, S.V. Morozov, L. Britnell, R. Jalil, L.A. (b) X. Qian, J. Liu, L. Fu, J. Li, Science 346 (2014) 1344–1347;
Ponomarenko, P. Blake, K.S. Novoselov, K. Watanabe, T. Taniguchi, A.K. Geim, (c) J. Wunderlich, B. Park, A.C. Irvine, L.P. Zarbo, E. Rozkotova, P. Nemec, V.
Nano Lett. 11 (2011) 2396–2399; Novak, J. Sinova, T. Jungwirth, Science 330 (2010) 1801;
(b) K.S. Novoselov, D. Jiang, F. Schedin, T.J. Booth, V.V. Khotkevich, S.V. (d) D. Pesin, A.H. MacDonald, Nat. Mater. 11 (2012) 409;
Morozov, A.K. Geim, Proc. Nat. Acad. Sci. U S A 102 (2005) 10451; (e) C.H. Li, O.M.J. van‘t Erve, J.T. Robinson, Y. Liu, L. Li, B.T. Jonker, Nat.
(c) D. Jariwala, V.K. Sangwan, D.J. Late, J.E. Johns, V.P. Dravid, T.J. Marks, L.J. Nanotechnol. 9 (2014) 218–224;
Lauhon, M.C. Hersam, Appl. Phys. Lett. 102 (2013) 173107; (f) L. Liu, J. Guo, J. Appl. Phys. 118 (2015) 124502;
(d) L. Liu, S.B. Kumar, Y. Ouyang, J. Guo, IEEE Trans. Electron Devices 58 (2011) (g) H. Yang, S.W. Kim, M. Chhowalla, Y.H. Lee, Nat. Phys. 13 (2017) 931.
3042–3047; [22] (a) Y. Sun, C. Felser, B. Yan, Phys. Rev. B 92 (2015) 165421;
(e) R. Fivaz, E. Mooser, Phys. Rev. 163 (1967) 743–755. (b) S.M. Nie, Z. Song, H. Weng, Z. Fang, Phys. Rev. B 91 (2015) 235434;
[12] (a) K. Lee, et al., Adv. Mater. 23 (2011) 4178–4182; (c) Y. Ma, L. Kou, X. Li, Y. Dai, S.C. Smith, T. Heine, Phys. Rev. B 92 (2015)
(b) W. Zhang, J.K. Huang, C.H. Chen, Y.H. Chang, Y.J. Cheng, L.J. Li, Adv. Mater. 085427;
25 (2013) 3456–3461; (d) Y. Ma, L. Kou, X. Li, Y. Dai, T. Heine, Phys. Rev. B 93 (2016) 035442;
(c) K. Kaasbjerg, K.S. Thygesen, K.W. Jacobsen, Phys. Rev. B 85 (2012) (e) A.M. van der Zande, P.Y. Huang, D.A. Chenet, T.C. Berkelbach, Y. You, G.H.
115317; Lee, T.F. Heinz, D.R. Reichman, D.A. Muller, J.C. Hone, Nat. Mater. 12 (2013)
(d) A.C. Ferrari, J.C. Meyer, V. Scardaci, C. Casiraghi, M. Lazzeri, F. Mauri, S. 554.
Piscanec, D. Jiang, K.S. Novoselov, S. Roth, A.K. Geim, Phys. Rev. Lett. 97 (2006) [23] (a) Y.L. Chen, J.G. Analytis, J.H. Chu, Z.K. Liu, S.K. Mo, X.L. Qi, H.J. Zhang, D.H. Lu,
187401–187404; X. Dai, Z. Fang, S.C. Zhang, I.R. Fisher, Z. Hussain, Z.X. Shen, Science 325 (2009)
(e) A. Gupta, G. Chen, P. Joshi, S. Tadigadapa, P.C. Eklund, Nano Lett. 6 (2006) 178;
2667–2673. (b) Y. Xu, B.H. Yan, H.J. Zhang, J. Wang, G. Xu, P. Tang, W.H. Duan, S.C. Zhang,
[13] (a) D. Graf, F. Molitor, K. Ensslin, C. Stampfer, A. Jungen, C. Hierold, L. Wirtz, Phys. Rev. Lett. 111 (2013) 136804;
Nano Lett. 7 (2007) 238–242; (c) Y.D. Ma, Y. Dai, L.Z. Kou, T. Frauenheim, T. Heine, Nano Lett. 15 (2015)
(b) C. Lee, H. Yan, L.E. Brus, T.F. Heinz, J. Hone, S. Ryu, ACS Nano 4 (2010) 1083;
2695–2700; (d) P.F. Liu, L. Zhou, T. Frauenheim, L.M. Wu, Nanoscale 9 (2016) 4915.
(c) H. Li, Q. Zhang, C.C.R. Yap, B.K. Tay, T.H.T. Edwin, A. Olivier, D. Baillargeat, [24] (a) M.Y. Li, Y. Shi, C.C. Cheng, L.S. Lu, Y.C. Lin, H.L. Tang, M.L. Tsai, C.W. Chu,
Adv. Funct. Mater. 22 (2012) 1385–1390; K.H. Wei, J.H. He, W.H. Chang, K. Suenaga, L.J. Li, Science 349 (2015) 524;
(d) B. Chakraborty, H.S.S.R. Matte, A.K. Sood, C.N.R. Rao, J. Raman Spectrosc. 44 (b) M.H. Chiu, C. Zhang, H.W. Shiu, C.P. Chuu, C.H. Chen, C.Y. Chang, C.H. Chen,
(2013) 92–96; M.Y. Chou, C.K. Shih, L.J. Li, Nat. Commun. 6 (2015) 7666;
(e) A.G. Bagnall, W.Y. Liang, E.A. Marseglia, B. Welber, Physica B 99 (1980) (c) M.H. Chiu, M.Y. Li, W. Zhang, W.T. Hsu, W.H. Chang, M. Terrones, H.
343–346. Terrones, L.J. Li, ACS Nano 8 (2014) 9649;
[14] (a) S. Bertolazzi, J. Brivio, A. Kis, ACS Nano 5 (2011) 9703–9709; (d) T.H. Ly, M.H. Chiu, M.Y. Li, J. Zhao, D.J. Perello, M.O. Cichocka, H.M. Oh, S.H.
(b) A.C. Gomez, M. Poot, G.A. Steele, H.S.J. Zant, N. Agraït, G.R. Bollinger, Adv. Chae, H.Y. Jeong, F. Yao, L.J. Li, Y.H. Lee, ACS Nano 8 (2014) 11401–11408;
Mater. 24 (2012) 772–775; (e) J.K. Huang, J. Pu, C.L. Hsu, M.H. Chiu, Z.Y. Juang, Y.H. Chang, W.H. Chang, Y.
(c) B. Radisavljevic, A. Radenovic, J. Brivio, V. Giacometti, A. Kis, Nat. Iwasa, T. Takenobu, L.J. Li, ACS Nano 8 (2014) 923.
Nanotechnol. 6 (2011) 147–150; [25] (a) C.H. Chen, C.L. Wu, J. Pu, M.H. Chiu, P. Kumar, T. Takenobu, L.J. Li, 2D Mater
(d) W. Zhao, Z. Ghorannevis, L. Chu, M. Toh, K. Kloc, P.H. Tan, G. Eda, ACS Nano 1 (2014) 034001;
7 (2013) 791–797. (b) Y.C. Lin, W. Zhang, J.K. Huang, K.K. Li, Y.H. Lee, C.T. Liang, C.W. Chu, L.J. Li,
[15] (a) R.M. Martin, Electronic structure Basic Theory and Practical Methods, Nanoscale 4 (2012) 6637–6641;
Cambridge University Press, 2004; (c) Y.H. Lee, X.Q. Zhang, W. Zhang, M.T. Chang, C.T. Lin, K.D. Chang, Y.C. Yu,
(b) J.G. Lee, Computational Materials Science: An introduction, CRC Press, J.T.W. Wang, C.S. Chang, L.J. Li, T.W. Lin, Adv. Mater. 24 (2012) 2320–2325;
New York, 2012; (d) G.H. Han, et al., Nano Lett. 11 (2011) 4144–4148;
(c) G. Kresse, J. Furthmuller, Phys. Rev. B 54 (1996) 11169; (e) J. Gao, J. Yip, J. Zhao, B.I. Yakobson, F. Ding, J. Am. Chem. Soc. 133 (2011)
(d) A. Bansil, H. Lin, T. Das, Reviews of Modern Physics 88 (2016) 021004. 5009–5015.
[16] (a) W. Kohn, Rev. Mod. Phys. 71 (1999) 1253; [26] (a) Y. Zhan, Z. Liu, S. Najmaei, P.M. Ajayan, J. Lou, Small 8 (2012) 966–971;
(b) R. Prasad, Electronic Structure of Materials, CRC Press, New York, 2014. (b) K.K. Liu, et al., Nano Lett. 12 (2012) 1538–1544;
270 A.K. Singh et al. / Applied Materials Today 13 (2018) 242–270

(c) Z. Guo, H. Zhang, S. Lu, Z. Wang, S. Tang, J. Shao, Z. Sun, H. Xie, H. Wang, [34] (a) M.S. Faber, R. Dziedzic, M.A. Lukowski, N.S. Kaiser, Q. Ding, S. Jin, J. Am.
X.-F. Yu, P.K. Chu, Adv. Funct. Mater. 25 (2015) 6996–7002; Chem. Soc. 136 (2014) 10053–10061;
(d) P. Joensen, R.F. Frindt, S.R. Morrison, Mater. Res. Bull. 21 (1986) 457–461; (b) M. Caban-Acevedo, M.L. Stone, J.R. Schmidt, J.G. Thomas, Q. Ding, H.C.
(e) W.M.R. Divigalpitiya, S.R. Morrison, R.F. Frindt, Thin Solid Films 186 (1990) Chang, M.L. Tsai, J.H. He, S. Jin, Nat. Mater. 14 (2015) 1245–1251;
177–192; (c) P.C.K. Vesborg, B. Seger, I. Chorkendorff, J. Phys. Chem. Lett. 6 (2015)
(f) A.H. Woomer, T.W. Farnsworth, J. Hu, R.A. Wells, C.L. Donley, S.C. Warren, 951–957;
ACS Nano 9 (2015) 8869–8884. (d) E.J. Popczun, J.R. McKone, C.G. Read, A.J. Biacchi, A.M. Wiltrout, N.S. Lewis,
[27] (a) W.M.R. Divigalpitiya, R.F. Frindt, S.R. Morrison, Science 246 (1989) R.E. Schaak, J. Am. Chem. Soc. 135 (2013) 9267–9270;
369–371; (f) J.F. Callejas, C.G. Read, C.W. Roske, N.S. Lewis, R.E. Schaak, Chem. Mater. 28
(b) S. Kirmayer, E. Aharon, E. Dovgolevsky, M. Kalina, G.L. Frey, Philos. Trans. (2016) 6027–6044.
R. Soc. A 365 (2007) 1489–1508; [35] (a) M. Chhowalla, H.S. Shin, G. Eda, L.J. Li, K.P. Loh, H. Zhang, Nat. Chem. 5
(c) G.L. Frey, K.J. Reynolds, R.H. Friend, H. Cohen, Y. Feldman, J. Am. Chem. Soc. (2013) 263–275;
125 (2003) 5998–6007; (b) M.S. Xu, T. Liang, M.M. Shi, H.Z. Chen, Chem. Rev. 113 (2013) 3766–3798;
(d) J. Heising, M.G. Kanatzidis, J. Am. Chem. Soc. 121 (1999) 638–643; (c) Q.H. Wang, K. Kalantar-Zadeh, A. Kis, J.N. Coleman, M.S. Strano, Nat. Nano
(e) M.A. Py, R.R. Haering, Can. J. Phys. 61 (1983) 76–84; 7 (2012) 699–712;
(f) M.A. Bissett, I.A. Kinloch, R.A.W. Dryfe, ACS Appl. Mater. Interfaces 7 (2015) (d) D. Merki, X.L. Hu, Energy Environ. Sci. 4 (2011) 3878–3888;
17388–17398; (e) H.I. Karunadasa, E. Montalvo, Y. Sun, M. Majda, J.R. Long, C.J. Chang,
(g) S. Wu, Z. Zeng, Q. He, Z. Wang, S.J. Wang, Y. Du, Z. Yin, X. Sun, W. Chen, H. Science 335 (2012) 698–702;
Zhang, Small 8 (2012) 2264–2270. (f) H. Tributsch, J.C. Bennett, J. Electroanal. Chem. 81 (1977) 97–111.
[28] (a) G. Eda, H. Yamaguchi, D. Voiry, T. Fujita, M.W. Chen, M. Chhowalla, Nano [36] (a) Y.G. Li, H.L. Wang, L.M. Xie, Y.Y. Liang, G.S. Hong, H.J. Dai, J. Am. Chem. Soc.
Lett. 11 (2011) 5111–5116; 133 (2011) 7296–7299;
(b) S. Hussain, J. Singh, D. Vikraman, A.K. Singh, M.Z. Iqbal, M.F. Khan, P. (b) T.F. Jaramillo, K.P. Jørgensen, J. Bonde, J.H. Nielsen, S. Horch, I.
Kumar, D.C. Choi, W. Song, K.S. An, J. Eom, W.G. Lee, J.W. Jung, Scientific Chorkendorff, Science 317 (2007) 100–102;
Reports 6 (2016) 30791; (c) J. Kibsgaard, Z.B. Chen, B.N. Reinecke, T.F. Jaramillo, Nat. Mater. 11 (2012)
(c) S. Hussain, M.A. Shehzad, D. Vikraman, M.F. Khan, J. Singh, D.C. Choi, Y. 963–969;
Seo, J. Eom, W.G. Lee, J.W. Jung, Nanoscale 8 (2016) 4340–4347; (d) J. Xie, H. Zhang, S. Li, R. Wang, X. Sun, M. Zhou, J. Zhou, X.W. Lou (David), Y.
(d) L.K. Tan, B. Liu, J.H. Teng, S. Guo, H.Y. Lowd, K.P. Loh, Nanoscale 6 (2014) Xie, Adv. Mater. 25 (2013) 5807–5813;
10584–10588; (e) D.Y. Chung, S.K. Park, Y.H. Chung, S.H. Yu, D.H. Lim, N. Jung, H.C. Ham, H.Y.
(e) B.D. Keller, A. Bertuch, J. Provine, G. Sundaram, N. Ferralis, J.C. Grossman, Park, Y. Piao, S.J. Yoo, Y.E. Sung, Nanoscale 6 (2014) 2131;
Chem. Mater. 29 (2017) 2024–2032. (f) G. Ye, Y. Gong, J. Lin, B. Li, Y. He, S.T. Pantelides, W. Zhou, R. Vajtai, P.M.
[29] (a) V. Podzorov, M.E. Gershenson, C. Kloc, R. Zeis, E. Bucher, Appl. Phys. Lett. Ajayan, Nano Lett. 16 (2016) 1097;
84 (2004) 3301–3303; (g) H. Li, C. Tsai, A.L. Koh, L. Cai, A.W. Contryman, A.H. Fragapane, J. Zhao, H.S.
(b) S. Walia, S. Balendhran, Y. Wang, A. Kadir, R. Zoolfakar, A. Atkin, P. Ou, J. Han, H.C. Manoharan, F. Abild-Pedersen, Nat. Mater. 15 (2016) 48;
Sriram, S. Kalantar, K. Zadeh, M. Bhaskaran, Appl. Phys. Lett. 103 (2013) (h) J. Luxa, V. Mazánek, D. Bouša, D. Sedmidubský, M. Pumera, Z. Sofer,
232105; ChemElectroChem 3 (2016) 565–571.
(c) J.H. Kang, W. Liu, K. Banerjee, Appl. Phys. Lett. 104 (2014) 093106; [37] (a) D.S. Kong, H.T. Wang, J.J. Cha, M. Pasta, K.J. Koski, J. Yao, Y. Cui, Nano Lett
(d) J. Yoon, W. Park, G.Y. Bae, Y. Kim, H.S. Jang, Y. Hyun, S.K. Lim, Y.H. Kahng, 13 (2013) 1341–1347;
W.K. Hong, B.H. Lee, Small (2019) 3295–3300; (b) M.A. Lukowski, A.S. Daniel, F. Meng, A. Forticaux, L.S. Li, S. Jin, J. Am. Chem.
(e) Y.T. Lee, K. Choi, H.S. Lee, S.W. Min, P.J. Jeon, D.K. Hwang, H.J. Choi, S. Im, Soc. 135 (2013) 10274–10277;
Small 10 (2014) 2356–2361; (c) L. Yang, H. Hong, Q. Fu, Y.F. Huang, J.Y. Zhang, X.D. Cui, Z.Y. Fan, K.H. Liu, B.
(f) S. Andleeb, J. Eom, N.R. Naz, A.K. Singh, J. Mater. Chem. C 5 (2017) 8308. Xiang, ACS Nano 9 (2015) 6478–6483;
[30] (a) S. Das, R. Gulotty, A.V. Sumant, A. Roelofs, Nano Lett. 14 (2014) 2861–2866; (d) L. Liao, J. Zhu, X.J. Bian, L.N. Zhu, M.D. Scanlon, H.H. Girault, B.H. Liu, Adv.
(b) T. Roy, M. Tosun, J.S. Kang, A.B. Sachid, S.B. Desai, M. Hettick, C.C. Hu, A. Funct. Mater. 23 (2013) 5326–5333.
Javey, ACS Nano 8 (2014) 6259–6264; [38] (a) J. Xie, J. Zhang, S. Li, F. Grote, X. Zhang, H. Zhang, R. Wang, Y. Lei, B. Pan, Y.
(c) A.K. Singh, C. Hwang, J. Eom, ACS Appl. Mater. Interfaces 8 (2016) Xie, J. Am. Chem. Soc. 135 (2013) 17881–17888;
34699–34705; (b) L. Tao, X. Duan, C. Wang, X. Duan, S. Wang, Chem. Commun. 51 (2015)
(d) A.K. Singh, S. Andleeb, J. Singh, J. Eom, RSC Adv. 5 (2015) 77014–77018; 7470–7473;
(e) A.K. Singh, R.K. Pandey, R. Prakash, J. Eom, Appl. Surf. Sci. 437 (2018) (c) G. Ye, Y. Gong, J. Lin, B. Li, Y. He, S.T. Pantelides, W. Zhou, R. Vajtai, P.M.
70–74; Ajayan, Nano Lett 16 (2016) 1097–1103;
(f) M. Shanmugam, C.A. Durcan, B. Yu, Nanoscale 4 (2012) 7399–7405. (d) D. Merki, S. Fierro, H. Vrubel, X.L. Hu, Chem. Sci. 2 (2011) 1262–1267;
[31] (a) S. Chu, A. Majumdar, Nature 488 (2012) 294–303; (e) H. Vrubel, X.L. Hu, ACS Catal. 3 (2013) 2002–2011.
(b) J.A. Turner, Science 305 (2004) 972–974; [39] (a) D. Merki, H. Vrubel, L. Rovelli, S. Fierro, X.L. Hu, Chem. Sci. 3 (2012)
(c) N.S. Lewis, D.G. Nocera, Proc. Natl. Acad. Sci. USA 103 (2006) 15729–15735; 2515–2525;
(d) M.G. Walter, E.L. Warren, J.R. McKone, S.W. Boettcher, Q.X. Mi, E.A. Santori, (b) J.D. Benck, Z.B. Chen, L.Y. Kuritzky, A.J. Forman, T.F. Jaramillo, ACS Catal. 2
N.S. Lewis, Chem. Rev. 110 (2010) 6446–6473; (2012) 1916–1923;
(e) A.J. Bard, M.A. Fox, Acc. Chem. Res. 28 (1995) 141–145; (c) H. Vrubel, D. Merki, X.L. Hu, Energy Environ. Sci. 5 (2012) 6136–6144;
(f) J.R. McKone, N.S. Lewis, H.B. Gray, Chem. Mater. 26 (2014) 407–414; (d) Y.F. Huang, R.J. Nielsen, W.A. Goddard, M.P. Soriaga, J. Am. Chem. Soc. 137
(g) A. Fujishima, K. Honda, Nature 238 (1972) 37–38. (2015) 6692–6698;
[32] (a) J.O.M. Bockris, E.C. Potter, J. Electrochem. Soc. 99 (1952) 169–186; (e) P.D. Tran, T.V. Tran, M. Orio, S. Torelli, Q.D. Truong, K. Nayuki, Y. Sasaki, S.Y.
(b) Y. Yan, X. Ge, Z. Liu, J.Y. Wang, J.M. Lee, X. Wang, Nanoscale 5 (2013) Chiam, R. Yi, I. Honma, et al., Nat. Mater. 15 (2016) 640.
7768–7771; [40] (a) Y. Yin, J. Han, Y. Zhang, X. Zhang, P. Xu, Q. Yuan, L. Samad, X. Wang, Y.
(c) X. Guo, G.I. Cao, F. Ding, X. Li, S. Zhen, Y. Xue, Y. Yan, T. Liu, K.J. Sun, Mater. Wang, Z. Zhang, et al., J. Am. Chem. Soc. 13 (2016) 7965–7972;
Chem. A 3 (2015) 5041–5046; (b) J. Zhang, J. Wu, H. Guo, W. Chen, J. Yuan, U. Martinez, G. Gupta, A.D.
(d) A.B. Laursen, A.S. Varela, F. Dionigi, H. Fanchiu, C. Miller, O.L. Trinhammer, Mohite, P.M. Ajayan, J. Lou, Adv. Mater. 29 (2017) 1701955.
J. Rossmeisl, S.J. Dahl, Chem. Educ. 89 (2012) 1595–1599; [41] (a) J.M. Velazquez, J. John, D.V. Esposito, A. Pieterick, R. Pala, G. Sun, X. Zhou, Z.
(e) J.K. Norskov, T. Bligaard, A. Logadottir, J.R. Kitchin, J.G. Chen, S. Pandelov, Huang, S. Ardo, M.P. Soriaga, et al., Energy Environ. Sci. 9 (2016) 164–175;
J.K. Norskov, J. Electrochem. Soc. 152 (2005) J23–J26; (b) Y. Qi, Q. Xu, Y. Wang, B. Yan, Y. Ren, Z. Chen, ACS Nano 10 (2016)
(f) B. Hinnemann, P.G. Moses, J. Bonde, K.P. Jorgensen, J.H. Nielsen, S. Horch, I. 2903–2909;
Chorkendorff, J.K. Norskov, J. Am. Chem. Soc. 127 (2005) 5308–5309; (c) M. Velický, M.A. Bissett, C.R. Woods, P.S. Toth, T. Georgiou, I.A. Kinloch, K.S.
(g) J. Greeley, T.F. Jaramillo, J. Bonde, I.B. Chorkendorff, J.K. Norskov, Nat. Novoselov, R.A.W. Dryfe, Nano Lett. 16 (2016) 2023–2032;
Mater. 5 (2006) 909–913. (d) J. Wong, D. Jariwala, G. Tagliabue, K. Tat, A.R. Davoyan, M.C. Sherrott, H.A.
[33] (a) M.S. Faber, S. Jin, Energy Environ. Sci. 7 (2014) 3519–3542; Atwater, ACS Nano. 11 (2017) 7230–7240;
(b) M. Zeng, Y.G. Li, J. Mater. Chem. A 3 (2015) 14942–14962; (e) D. Jariwala, A.R. Davoyan, G. Tagliabue, M.C. Sherrott, J. Wong, H.A.
(c) J.D. Benck, T.R. Hellstern, J. Kibsgaard, P. Chakthranont, T.F. Jaramillo, ACS Atwater, Nano Lett. 16 (2016) 5482–5487.
Catal. 4 (2014) 3957–3971; [42] (a) Q. Ding, F. Meng, C.R. English, M. Caban-Acevedo, M.J. Shearer, D. Liang,
(d) A.B. Laursen, S. Kegnaes, S. Dahl, I. Chorkendorff, Energy Environ. Sci. 5 A.S. Daniel, R.J. Hamers, S. Jin, J. Am. Chem. Soc. 136 (2014) 8504–8507;
(2012) 5577–5591; (b) K. Chang, M. Li, T. Wang, S. Ouyang, P. Li, L. Liu, J. Ye, Adv. Energy Mater. 5
(e) C.G. Morales-Guio, X.L. Hu, Acc. Chem. Res. 47 (2014) 2671–2681; (2015);
(f) S. Jayabal, G. Saranya, J. Wu, Y. Liu, D. Geng, X. Meng, J. Mater. Chem. A 5 (c) K. Chang, Z. Mei, T. Wang, Q. Kang, S. Ouyang, J. Ye, ACS Nano. 8 (2014)
(2017) 24540–24563; 7078–7087;
(g) A. Eftekhari, Appl. Mater. Today 8 (2017) 1–17; (d) E. Ha, W. Liu, L. Wang, H.W. Man, L. Hu, S.C.E. Tsang, C.T.L. Chan, W.M.
(h) M. Velický, P.S. Toth, Appl. Mater. Today 8 (2017) 68–103; Kwok, L.Y.S. Lee, K.Y. Wong, Sci. Reports 7 (2017) 39411.
(i) L. Yang, P. Liu, J. Li, B. Xiang, Catalysts 7 (2017) 285, http://dx.doi.org/10.
3390/catal7100285.

You might also like