0% found this document useful (0 votes)
7 views76 pages

Photodetectors Based On Two-Dimensional Layered Materials Beyond Graphene

Uploaded by

hugmelrose
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)
7 views76 pages

Photodetectors Based On Two-Dimensional Layered Materials Beyond Graphene

Uploaded by

hugmelrose
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/ 76

This is the Pre-Published Version.

This is the peer reviewed version of the following article: Xie, C., Mak, C., Tao, X., & Yan, F. (2017). Photodetectors Based on Two-Dimensional
Layered Materials Beyond Graphene. Advanced Functional Materials, 27(19), 1603886, which has been published in final form at
https://doi.org/10.1002/adfm.201603886. This article may be used for non-commercial purposes in accordance with Wiley Terms and Conditions for
Use of Self-Archived Versions.
Photodetectors Based on Two-Dimensional Layered
Materials Beyond Graphene
Chao Xie, Chunhin Mak, Xiaoming Tao, and Feng Yan*

Dr. C. Xie, C. H. Mak, Prof. F. Yan Department of Applied Physics

The Hong Kong Polytechnic University Hong Kong, China

E-mail: [email protected]

Prof. X. M. Tao Institute of Textiles & Clothing

The Hong Kong Polytechnic University Hong Kong, China

Abstract
Following a significant number of graphene studies, other two-dimensional (2D)
layered materials have attracted more and more interest for their unique structures
and distinct physical properties, which has opened a window for realizing novel
electronic or optoelectronic devices. Here, we present a comprehensive review on
the applications of 2D-layered semiconductors as photodetectors, including
photoconductors, phototransistors, and photodiodes, reported in the past five years.
The device designs, mechanisms, and performances of the photodetectors are
introduced and discussed systematically. Emerging techniques to improve device
performances by enhancing light-matter interactions are addressed as well. Finally,
we deliver a summary and outlook to provide a guideline of the future development
of this rapidly growing field.
1. Introduction
Two-dimensional (2D) layered materials refer to a class of mate- rials artificially
derived from layered van der Waals solids,[1,2] in which the atoms are held together
by tight covalent or ionic bonds along 2D (in-plane) directions to form atomic
layers, while the atomic layers are bonded together by weak van der Waals
interactions along the third-dimensional (out-of-plane) direction.[3,4] Such weak van
der Waals interactions between neighboring layers make it possible to cleave the
layered mate- rials into individual freestanding few-atom thick—or even single-
atom thick—layers via mechanical[5,6] or liquid phase exfoliation[7,8] techniques.
Discovered in 2004, graphene, composed of a single layer of carbon atoms bonded
together in a hexagonal honeycomb lattice, is the most famous 2D layered
material[9] and has plenty of appealing electronic, optical, mechanical, and thermal
properties.[10–12] For example, graphene is gapless and semi-metallic, allowing it to
interact with light over a broad bandwidth from terahertz (THz) to ultra-violet (UV)
wavelengths[13,14] and rendering it a promising material for various photodetectors
over a wide spectral range.
However, despite its high absorption coefficient,[15] single-layer graphene can
absorb only 2.3% of incident visible and infrared (IR) light due to its thinness, [16,17]
which is undesirable for high performance photodetectors that require strong light
absorption. Graphene’s gaplessness also leads to short photocarrier lifetime in pure

This article may not be enhanced, enriched or otherwise transformed into a derivative work, without express permission from Wiley or by statutory rights
under applicable legislation. Copyright notices must not be removed, obscured or modified. The article must be linked to Wiley’s version of record on
Wiley Online Library and any embedding, framing or otherwise making available the article or pages thereof by third parties from platforms, services and
websites other than Wiley Online Library must be prohibited.
graphene, which is unfavorable for efficient photocurrent generation. [18–20] In
addition to graphene, this family of materials contains an extremely large range of
other 2D-layered materials that have recently attracted much research interest.[1-8]
Similar to graphene, being atomically thin, these materials exhibit a wide range of
unique electrical, optical, thermal, and mechanical properties that can never be seen
in their three-dimensional (3D) bulk counterparts due to the dimensionality
confinement effect and modulation in their band structures.[21,22] These 2D-layered
materials can be metals, semiconductors, insulators, superconductors, topological
insulators, or paramagnetic, diamagnetic, ferromagnetic, or anti-ferromagnetic,
etc., depending on their composition and phases.[23,24] Among these materials,
particular attention has been paid to 2D layered semiconductors thanks to their
unique electronic[2,24–26] and optoelectronic properties,[27,28] which arise from their
appreciable bandgaps ranging from IR to UV and throughout the visible range.
The optical and optoelectronic properties of 2D layered semiconductors are strongly
dependent on their number of layers due to the quantum confinement effects in the
out-of-plane direction and changes in symmetry.[29,30] This layer-dependent
modulation of physical properties, such as bandgap structure, is particularly
evident in the semiconducting transition metal dichalcogenides (TMDs)[2,7,29,31,32]
and few-layer black phosphorus (BP).[33,34] Another effect induced by vertical
quantum confinement is the increased absorption efficiency, which results from the
strongly bound excitons due to the reduced thick- ness.[29,35,36] The atomic thickness
of a 2D layered semiconductor also leads to high transparency[37] and good
mechanical flexibility,[38] properties of particular interest for novel device applications
such as flexible, wearable, or portable electronics. Despite their high transparency,
2D layered semiconductors can strongly interact with incident light, leading to
enhanced photon absorption and electron–hole creation due to the existence of Van
Hove singularities in their electronic density of states.[39] It has also been reported
that 2D layered materials possess extraordinary elastic modulus and large strain
(>10%) before rupture,[40–43] which allow for tuning their electronic and optical
properties via strain engineering and the implementation of novel device
architectures such as wearable, bendable devices and devices with novel
functionalities. Moreover, the van der Waals inter- actions between neighboring
layers without surface dangling bonds enable the integration of materials with
properties beyond the limitation of lattice matching [44,45] or their direct growth on
a variety of substrates. [46,47] This paves the way for designing various functional
heterostructures with fundamentally different properties and understanding the
underlying mechanisms. Further band engineering of these heterostructures is
achievable due to the diversity of 2D layered materials with different bandgaps and
work functions. In addition, the ultimate thickness of 2D layered materials offers
us the ability to tune their carrier densities, band alignments, and polarities via
electrostatic [25,48,49] or chemical [50–52] doping, providing an alternative way to
construct functional heterostructures. The wide range of tunable material
properties, along with the possibilities for creating various functional
heterostructures, render 2D-layered semiconductors promising candidates for future
high-performance optoelectronic applications, including various photodetectors
covering a wide spectral range from IR to UV wavelengths.
Photodetectors that can convert light signals into electrical signals are significant
in the development of a multitude of innovative technologies in modern society
from a practical perspective. These technologies include video and biomedical
imaging, optical communication, environment monitoring, military, security
checks, night vision, scientific research, and industrial processing control, among
others.[15–17,37] Currently, high-performance photodetectors made from crystalline
silicon (Si) with a detection range at visible and near- infrared (NIR) dominate the
photodetection market, while longer-wavelength photodetectors beyond the
detection limit of Si-based products are primarily constructed with materials such
as InGaAs and related heterostructures. However, these detectors usually suffer
from severe drawbacks, including thick material usage, expensive fabrication
processes, strictly controlled fabrication conditions, and rigorous operation
demands. These materials in bulk form are also very brittle, making the detectors
unsuitable for some novel device concepts, such as transparent, flexible, and
bendable applications. Therefore, there is an increasing demand for new
materials that can provide high absorption efficiencies in broadband wavelengths,
high transparency and flexibility, and can be processed using cost-effective
techniques with simplified device geometries to overcome the limitation of bulk Si
for photo- detection. The emergence of 2D layered semiconductors with appealing
properties such as high transparency, strong light- matter interactions, good
flexibility, ease of processing, etc., complements the shortages of current Si
technologies and affords great promise towards the realization of high-performance
photodetectors.
In this paper, we comprehensively review the photodetection applications of 2D
layered semiconductors beyond graphene in recent years, as depicted in Figure
1. Photodetectors based on graphene have already been reviewed in several papers
and are thus not covered in detail in this paper. [15–17] We first give a brief
introduction of the most commonly used synthetic methods for preparing 2D
layered materials along with the basic sensing mechanisms and important
performance parameters in photodetection. We then review photoconductors and
phototransistors based on a variety of 2D layered semiconductors and their hybrids.
Next, photodiodes, another type of photodetector based on these 2D layered
semiconductors, and their artificial heterostructures are discussed. Some emerging
techniques for light-matter inter- action enhancement to improve device
performance are also addressed. Finally, the existing techniques are summarized
and outlooks are provided to shed light on the future development of this research
area.
2.Device Fabrication and Physics
2.1. Synthesis Methods for 2D Layered Materials
The preparation of high-quality 2D layered materials with scalable and cost-
effective approaches is of significant importance for the realization of high-
performance photodetectors based on 2D layered materials. Recently, many
methods have been developed for synthesizing 2D layered materials, which can be

FEATURE ARTICLE
generally categorized into two types: top-down methods and bottom-up methods.[21,22]
Top-down methods, including techniques such as mechanical exfoliation, liquid
phase exfoliation, and so forth, refer to the synthesis of 2D layered materials by
breaking apart a single or few atomic layers from the bulk layered materials with the
aid of external energy (mechanical or ultrasound energy). In contrast, bottom-up
methods are those where atoms or molecules are assembled on a substrate via
thermal and/or chemical reactions for the synthesis of 2D layered materials.
Processes such as chemical vapor deposition, van der Waals epitaxial growth, and
hydrothermal synthesis fall into this category. In this section, a brief description of
the most commonly used synthetic methods for the production of 2D layered
materials will be presented, and the advantages and limitations of each approach
will be discussed. For a more detailed description of the preparation of 2D layered
materials, interested readers should refer to other related review papers. [21,22,26,46,53]

Figure 1. Timeline showing the development of the applications of graphene and other 2D
layered semiconductors in photodetectors based on different principles. Reproduced with
permission.[12,77,88,129,151,152,222,238,246,253,280,286,305,313] Copyright 2008, 2012, 2013, 2014,
2015, Nature Publishing Group. Copyright 2012, 2013, American Chemical Society.
Copyright 2013, 2015, 2016, Wiley-VCH.
2.1.1. Mechanical Exfoliation
The mechanical exfoliation method, invented by A. Geim and K. Novoselov to
cleave graphene from highly oriented pyrolytic graphite in 2004, is a traditional
technique for obtaining atomical and few-layer thick flakes of 2D layered inorganic
materials.[9] Following the isolation of graphene, this approach has been successful
for obtaining single or few layers of other non- carbonaceous 2D layered materials,
such as TMDs (MoS2,[2] MoSe2,[54] WS2,[55] WSe2,[56] etc.), h-BN[57] (hexagonal
boron nitride) and few-layer BP.[25] In a typical process, Scotch tape is used to
repeatedly stick and peel apart layered solid flakes into thinner layers until single-
or few-layer flakes are attained on the tape. The flakes are washed with acetone
and transferred onto an oxidized Si substrate. The number of layers in the exfoliated
flakes can be determined by optical microscopy, atomic force microscopy (AFM)
height profiles, scanning tunneling microscopy (STM), and high-resolution
transmission electron microscopy (HRTEM). The obtained 2D flakes, which
maintain the crystal structure of the bulk form, can remain stable in ambient
conditions for days to months. Mechanical exfoliation is a low cost and convenient
method for 2D layered materials production and is highly suitable for fundamental
research where 2D flakes with structural integrity, high crystallinity, minimal
defects are desired. However, scalable production is highly challenging due to the
low throughput from this process. In addition, this method lacks controllability in
uniformity, size, and thickness of the peeled flakes, and thus does not meet the
requirements for practical applications.
2.1.2. Liquid Phase Exfoliation
An alternative method to obtain single- or few-layer 2D mate rials is based on
liquid-phase exfoliation, which can be divided into solvent exfoliation and chemical
exfoliation. Solvent exfoliation generally involves the processes in which bulk layered
crystals or powders are dispersed in a suitable organic solvent and sonication is
employed as a source of external energy to cleave the bulk materials into atomically
thin sheets. Afterwards, the resultant dispersion is centrifuged to separate large
crystals; critical parameters that dictate the quality of the dispersion are sonication time
and centrifugation rate. Also, the selection of an appropriate solvent is another key
factor, and it has been suggested that solvents with surface tension similar to the
surface energy of 2D layered materials can not only minimize the energetic cost of
exfoliation but also facilitate the sheet separation.[21,46] A variety of 2D layered
materials such as MoS2, WS2, MoSe2, MoTe2, NbSe2, TaSe2, NiTe2, VS2, and h-BN
have been successfully synthesized using this method in some common solvents,
including NMP (N-methyl-2-pyrrolidone), DMF (dimethylformamide), DMSO
(dimethyl sulfoxide), IPA (isopropyl alcohol), cyclohexyl-pyrrolidinone, Ndodecyl-
pyrrolidone, NVP.[8,58–62] Usually, this method yields single- and few- layer 2D
nanosheets with micrometer dimensions dispersed homogeneously in a solvent. A
density gradient ultracentrifugation method can be used to separate single layers from
few- layers, which has proved to be an effective means of separating single-layer
graphene oxide from the solution.[63]
However, despite the low-cost, large-scale production ability of this method, it
faces drawbacks such as poor uniformity in layer thickness and low yields of large
flakes and single-layers, which are highly desired for electronic and optoelectronic
applications. To address this issue, a chemical exfoliation method has been
developed. In a typical process, the interlayers of a bulk layered material are
intercalated with small ions such as lithium ions, which break the interlayer van der
Waals bonding with external agitation and separate the layers. [21,54] A suspension
of single layers is readily obtained after washing with a suitable solvent to remove
excess lithium and other organic residues. This method has been successfully
applied for obtaining single layers of 2D TMDs, such as MoS [7,64] and WS2.[65]
Lithium-based intercalation exfoliation is a highly desirable and versatile process
that facilitates the low cost and large scale production of single layers of 2D
materials. However, some challenges still exist, such as homogeneous dispersion of
2D single layers and precise control of the layer size. Moreover, residual lithium
ions physically absorbed at the surface of the single layers are inevitably
incorporated in a lithium-based intercalation process, which deteriorate the
electrical and thermal properties of the final products. Another reported drawback
is the phase transition in MoS2 caused by the lithium intercalation, which can be
reversed by performing an additional thermal annealing process.[7]
Furthermore, with the aid of electrochemical systems, Zhang’s group has
developed an electrochemical lithium-based intercalation and exfoliation
method.[66,67] Briefly, they used a bulk layered crystal as cathode and a lithium foil
anode in a lithium ion battery setup. During the discharge process, lithium ions are
intercalated into the interlayers of the crystal. The lithium reacts with water or
ethanol to generate H2 gas, which facilitates the separation of adjacent layers.
Finally, nanosheets hundreds of nanometers in size, homogeneously dispersed in
a solvent, are obtained under vigorous agitation by sonication. Various 2D layered
materials, including MoS2, WS2, TiS2, TaS2, ZrS2, NbSe2, WSe2, Sb2Se3, Bi2Te3, h-
BN, and graphene single- layers have been successfully prepared by this method.
[66,67]
Compared with common lithium intercalation, this method can provide good
control over the degree of lithium intercalation by monitoring the discharge curve.
This means the amount of inserted lithium can be optimized to avoid insufficient or
excessive lithium intercalation.
2.1.3. Chemical Vapor Deposition
The chemical vapor deposition (CVD) technique has been extensively employed to
synthesize large-area homogeneous 2D TMDs with controllable thickness and
excellent proper- ties. Compared with the above discussed top-down methods, this
bottom-up method has many distinct advantages. In addition to the scalable size
and controllable thickness of the final products, the growth instruments are
inexpensive and versatile. Moreover, the preparation process is compatible with
current semiconductor technology. As an example, we describe the growth of 2D
MoS2 to introduce the synthesis of 2D TMDs via the CVD method. The most
commonly used CVD method involves MoO3 and sulfur powders as starting
materials to deposit single- or few-layer MoS2 on an SiO2/Si substrate. [47,68] At
high temperature under an inert atmosphere, MoO3 powders are reduced by the
sulfur vapor to form volatile suboxide MoO3–x, which diffuses to the substrate and
reacts with sulfurvapor to grow MoS2 layers with lateral sizes up to 2 mm. The
substrate can also be pretreated with graphene-like molecules, such as reduced
graphene oxide (rGO), perylene-3,4,9,10-tetracarboxylic acid tetrapotassium salt
(PTAS), and per- ylene-3,4,9,10-tetracarboxylic dianhydride (PTCDA) to promote the
layer growth of MoS2.[47] This method has also been applied to the growth of MoS2 or
WS2 layers on other insulating substrates including quartz and sapphire.[69] In
addition, wafer-scale MoS2 thin layers have been synthesized using MoO3 thin
films pre-deposited on a sapphire substrate as a starting material. This is followed by a
two-step thermal process, where reduction of MoO3 occurs in a H2/Ar environment and
2
further
. sulfurization in a sulfur-rich environment at elevated temperature yields MoS2
[70]
layers. In the second method, pre-deposited Mo thin layers on a SiO2/Si substrate
have been employed to react with sulfur vapor, which gives rise to large-area single-
or few-layer MoS [71] This process is highly scalable because the size and thickness
of the resultant MoS2 layers can be directly determined by the size and thickness of
the starting Mo films. In addition, Liu et al. reported that the thermal decomposition
of (NH4)2MoS4 layers dip-coated on various insulting substrates in the presence of
sulfur vapor via a two-step annealing process can produce large-area MoS2 layers,
providing an alternative strategy for synthesizing MoS2 with scalable size.[72]
The direct growth of MoS2 layers on insulting substrates can afford the distinct
advantage of skipping the film transfer step that is needed for CVD-growth
graphene. Thus, some drawbacks usually associated with this step, such as chemical
residues, possible physical film damage, and wrinkle formation can be avoided.
Also, MoS2 films produced by these methods are transferable to arbitrary substrates
with the help of PMMA (poly (methyl methacrylate)) and NaOH.[22]
Recently, four-inch wafer-scale films of monolayer MoS2 and WS2 with excellent
spatial homogeneity and coverage have been successfully grown on SiO2/Si
substrates using a metal- organic CVD technique.[73] In this approach, Mo(CO)6,
W(CO)6, (C2H5)2, and H2 are employed as gas-phase precursors, and argon is used
as the carrier gas. The resultant MoS2 layers show good electrical properties such
as a high electron mobility of 30 cm2V-1s-1, which is comparable to the
mechanically-exfoliated monolayer MoS2. Also, the wafer-scale batch fabrication
of monolayer MoS2 field effect transistors (FETs) has achieved a device yield of
99% with the use of these MoS2 films. These results are promising and show
potential towards the realization of wafer-scale electronics and optoelectronics
based on 2D TMDs.
2.2. Sensing Mechanisms
A photodetector is an optoelectronic device that can convert absorbed optical
energy into electrical signal, such as photocur rent or photovoltage. Several sensing
mechanisms have been reported, including photoconductive effect, photogating
effect, photovoltaic effect, photothermoelectric effect, and bolometric effect.
Based on the sensing mechanism, photodetectors can be divided into two
categories: photon (or quantum) detectors (including
photoconductors/phototransistors and photodiodes) and thermal detectors. In the
following section, we introduce the mechanisms for photodetectors based on 2D
layered semiconductors
2.2.1. Photoconductive Effect
The photoconductive effect involves a process in which photon absorption by a t
semiconductor generates excess free carriers and results in an increase in
conductivity. [74,75] A typical photoconductor consists of a semiconductor as a
channel, with two ohmic contacts affixed to opposite ends of the channel that
serve as source/drain electrodes. In darkness, the channel allows only a small
source-drain current to flow (dark current, Id) under a bias applied between the
electrodes (VDS). Under illumination, photons with energy larger than the bandgap
of the semiconductor are absorbed, which leads to the generation of electron–hole
pairs (or excitons depending on the material). The photoexcited electron-hole pairs
are then separated by the VDS, and the free electrons and holes drift oppositely
towards the source/drain electrodes. This process leads to a net increase
(photocurrent, Iph) in the channel current. It is noteworthy that carrier
recombination usually takes place at the surface and bulk of the semiconductor,
which leads to a decrease in photocurrent. Normally, a strong VDS is desired to
facilitate electron-hole pair separation and consequently carrier transportation. The
lifetime of the photocarriers has a significant effect on the photoconductive gain
(G) and the response speed. A photoconductor can have gain larger than unity.
Under a moderate VDS, the majority of carriers (assumed to be electrons) have a
much higher mobility and consequently a much shorter transit time than that of
the minority carriers (holes). Thus, the photoexcited electrons can drift across the
channel much more quickly than the photoexcited holes. To maintain charge
neutrality, more electrons are supplied from the other electrode and circulate many
times in the channel during the lifetime of holes. The gain depends upon the ratio
of lifetime of holes (τlifetime) to the electron transit time (τtransit) and is expressed
!"#$%&#'%
as G= . In order to obtain a high gain, the electron transit time has to be
!&()*+#&
short, while the lifetime of holes should be long. On the other hand, the response
time, which is related to the carrier recombination processes, is also determined by
the lifetime of holes. This means that high gain will reduce the response speed.
Therefore, a trade off has to be made between gain and response speed to get a
reasonable overall performance of a photoconductor.
2.2.2. Photogating Effect
The photogating effect can be regarded as a special case of the photoconductive
effect. Under illumination, electron-hole pairs are generated in the semiconductor,
and subsequently one of the two carrier types (electrons or holes) is trapped
in the localized states located at defects and/ or at surface adsorbates.[76]
Another case of electron–hole generation takes place in the surface adsorbates or
charge traps adjacent to the semiconductor serving as a conducting channel.[77,78]
Consequently, one type of carrier is transferred to the channel, leaving oppositely-
FEATURE ARTICLE

charged carriers trapped in the surface adsorbates or traps. These trapped carriers
act as local gates and effectively modulate the conductance of the semiconductor
due to electrostatic interactions. The carriers in the semiconductor can recirculate
m a n y times during the lifetime of the trapped carriers, leading to a high gain.
Notably, this effect is particularly pronounced in nanostructured materials with
large surface-area-to-volume ratios, such as colloidal quantum dots, nanowires, and
2D layered materials. Although photodetectors with a photogating effect can have
a higher gain than photoconductors, they usually suffer from slower response
speed. Sometimes, the two effects can take place in the same device and
contribute to the photocurrent simultaneously.
2.2.3. Photovoltaic Effect
The photovoltaic effect relies on a built-in electric field to separate the
photogenerated electron–hole pairs and propel electrons and holes towards
opposite directions.[79] The built- in electric field is normally produced at a junction
(a depleted semiconductor region), where there is a significant difference in the
work functions between materials. Photodetectors that opeate via the photovoltaic
effect are usually called photodiodes. Generally, the photodiode family contains p-
type/n-type (PN) photodiodes formed by two semiconductors with opposite doping
type and Schottky barrier photodiodes formed at the interface between a
semiconductor and a metal. A PIN photodiode is one example of a PN photodiode,
where “I” represents an intrinsic semiconductor layer inserted between the PN
junction. A photodiode usually displays an asymmetric current–voltage
characteristic (rectifying behavior) in the dark, while the device can function at two
modes under illumination, i.e., the photovoltaic mode (at zero bias) and the photo-
conductive mode (under reverse bias). In photovoltaic mode, the photogenerated
electron–hole pairs are separated by the built-in electric field and electrons and
holes are collected at opposite electrodes, which generates a considerable
photocurrent (short-circuit current, ISC). The electrical output can also be a
photovoltage (open-circuit voltage, VOC), which is obtained by keeping the circuit
open. A photodiode working in this mode has the lowest dark current, thus leading
to an improved detectivity, as well as maximized linearity and sensitivity. In
photoconductive mode, the external electric field with a moderate biasing voltage
has the same direction as the built-in one, which increases the separation
efficiency of the electron–hole pairs, as well as the response speed due to reduced
carrier transit time and lowered diode capacitance. In practice, a photodiode in
photovoltaic mode is well suited for precision applications, while a photodiode in
photoconductive mode is better suited for high-speed applications. Normally, a
photodiode has a maximum gain of unity, much smaller than a photodetector
operating via the photoconductive or photogating effect. However, too-large
reverse bias can cause avalanche multiplication or breakdown of a photodiode
(avalanche photodiode, APD), where photogenerated electrons with enough energy
initiate impact ionization, providing large internal current gain.
2.2.4. Photo-Thermoelectric Effect
The photo-thermoelectric effect occurs when non-uniform light-induced heating on
two dissimilar conductors or semiconductors leads to a temperature gradient,
causing a temperature difference (ΔT) between the two substances. This
temperature difference can result in a voltage difference, namely the photo-
thermoelectric voltage (VPTE), through the Seebeck effect (or thermoelectric effect).
The photo-thermoelectric voltage VPTE can be expressed as [80,81]
VPTE = (S1 −S2 )ΔT (1)
where S1, S2 are the Seebeck coefficients of the two substances. The photo-
thermoelectric voltage can produce a current through the circuit at zero applied bias.
Note that the photo-thermoelectric effect cannot be found in a uniform
semiconductor because negligible temperature gradients can be achieved.
Normally, the temperature gradient can be obtained by localized illuminating of a
measured device, or by globally illuminating a device with distinct absorption
coefficients in different parts.
2.2.5. Bolometric Effect
The bolometric effect is based on the resistance change of a material induced by
uniform heating under illumination.[82] Two factors, the conductance variation of a
material with temperature (dG/dT) and the homogeneous temperature increase (ΔT)
induced by the heating, determine the magnitude of this effect. Bolometers have
been widely used in astronomy, where sensitive detection of mid- and far-IR as
well as THz is much desired.[83] The bolometric effect has also been observed in
graphene-based photodetectors, where carriers can easily be heated due to weak
carrier-phonon coupling, [84,85] and in photo- detectors based on few-layer BP
with≈100 nm thickness.[86]
2.3. Photodetector Performance Parameters
In general, several important parameters are used to compare the detection
performance of photodetectors that operate under different mechanisms and are
constructed from different mate- rials and geometries. In the following section, we
will give a brief introduction to these parameters.
Responsivity (R): The responsivity R of a photodetector is defined as the ratio of
the output photocurrent or photovoltage to the input optical power on the active
region of the device, which is usually expressed as
,-.&./0((%*& 2 .( 3-.&.4."&)5% (7 ) <!" .( =!"
R= 9*30& .3&#/)" 3.:%( (;)
= >#$
(2)

where Iph is the photocurrent, Vph is the photovoltage, and Pin is the input optical
power. This parameter is used to indicate the available output photocurrent or
photovoltage of the photodetector for a given incident optical power at a certain
wavelength.
External Quantum Efficiency (EQE): The external quantum efficiency EQE is the
ratio between the number of electron-hole pairs with contribution to the
photocurrent and the number of incident photons. It can be expressed as
< ⁄? AC
EQE=> !"⁄AB=R?D (3)
#$

where e is the elementary charge, h is Planck’s constant, 𝑣 is the frequency of


incident light, c is the speed of incident light, and 𝜆 is the wavelength of incident
light. To achieve a large EQE in a photodetector, the optical absorption of the active
layer should be high, while the carrier recombination and trapping before being
collected should be minimized.
Internal Quantum Efficiency (IQE): Similarly, the internal quantum efficiency IQE
is the ratio between the number of electron-hole pairs with contribution to the
photocurrent and the number of absorbed photons by the detector. The number of
absorbed photons can be determined by deducting the photon losses, including
those due to transmission and reflection, from the total number of incident
photons. The optical interference effects should especially be taken into account
in photodetectors with an ultrathin active layer.
Speed and Bandwidth (Δf): The response speed determines the ability of a
photodetector to follow a rapid modulated optical signal. In the time domain, the
rise time tr and the fall time tf are typically used to characterize the response speed
of a photodetector. In some examples from the literature, the rise time tr and the
(fall time tf ) is defined as the time interval required for the the response to rise
(decay) from 10% (90%) to 90% (10%) of its peak value. A photodetector with a
fast response speed is highly desirable in certain areas, such as optical
communication, video imaging, etc.
Generally, the overall response speed of a photodetector is determined by its
bandwidth Δf, including the intrinsic bandwidth and radio frequency (RC) circuit-
limited bandwidth. The bandwidth of a photodetector is defined as the difference
between the upper (f2) and lower (f1) light modulation frequencies, where the
detector has uniform response in this range and no response outside of this range,
expressed as Δf= f2 – f1. Another widely used parameter in the frequency domain is
3-dB bandwidth (f3dB), which is the frequency at which the response drops to -3 dB
(50%) of its peak value and is related to Δf by f =0.886Δf. The f3dB of a photodetector
and its corresponding tr are related as follows:[87]
E.GH
tr=I (4)
%&'

Signal to Noise Ratio (SNR): Low noise is a significant factor for a photodetector
because it ultimately determines the lowest detectable signal strength. Noise always
exists in a photodetection process, which limits the detection of small amounts of
radiation energy by producing a random fluctuation in the output of the device. The
SNR of a photodetector is expressed as
J#5*)" 3.:%(
SNR = K.#+% 3.:%(
(5)

The SNR must be larger than 1 so that the signal power can be distinguished from
the noise.
Noise Equivalent Power (NEP): Noise equivalent power NEP is the minimum
impinging optical power required to achieve an SNR of 1 in a 1 Hz bandwidth.
FEATURE ARTICLE

The NEP is defined as


>(
NEP = (6)
LMI

where P1 is the incident optical power that produces SNR = 1.


The NEP can be written in some cases as
(/)
)
N$
𝑁𝐸𝑃 = O
(7)
n
R/Q
where 𝑖PQ is the mean-square noise current measured at the bandwidth of 1 Hz
in darkness (in A Hz–1/2). For a photodetector that has an output voltage signal,
(/)
)
B$ R/Q
the NEP= O
, where 𝑣PQ is the mean-square noise voltage
measured at the bandwidth of 1 Hz in darkness (in V Hz–1/2).
Detectivity (D*): The detectivity D* is a useful parameter for comparing the
detection performance of photodetectors with different materials and geometries,
and is defined as
(TMI)(/)
D* = UV>
(8)

where A is the device area and f is its bandwidth. Since the total device noise is
normally proportional to the square root of the area, the D* is independent of
the area. This parameter depends on sensitivity, spectral response, and noise of the
device. If the shot noise from dark current dominates the total noise of a
photodetector, the D* can be written as
T (/) O
D* ≈ (Q?< (/)
(9)
&)

Where 𝐼W is the dark current.


3. Photoconductors/Phototransistors Based on 2D Layered Semiconductors
A photoconductor/phototransistor is the simplest type of photo- detector, usually
operating with the photoconductive effect, photogating effect, or two effects
combined. A photoconductor can provide a high photocurrent gain at the expense
of reduced response speed. In particular, by exploiting the inherent amplification
function, a phototransistor can obtain an even higher gain. The electronic band
structure of a semiconductor greatly influences its optical absorption, which in turn
affects its opto- electronic properties. Due to the atomic thickness and large surface
area, photodetectors based on 2D layered semiconductors are expected to
demonstrate some novel phenomena and achieve superior performance compared
to their bulk or thin- film counterparts. In this section, we review the achievements
involving photoconductors/phototransistors based on 2D layered semiconductors.
3.1 Transition Metal Dichalcogenides (TMDs)
TMDs represent a class of layered materials with generalized formula MX2, where
M is a transition metal from group 4–7 and X is a chalcogen atom such as S, Se,
or Te.[22,24] Each layer of TMDs is composed of one layer of hexagonally packed
metal atoms with two atomic layers of chalcogen atoms covalently bonded on either
side. Similar to graphene, several layers are bonded together by weak van der Waals
interactions to form a bulk crystal. This weak interaction along the out-of-plane
direction allows the bulk crystals to be isolated down to single layers, which offers
the opportunity for fundamental research and proof-of-concept device fabrication
based on 2D TMDs. Single- or few-layer semiconducting TMDs hold unique optical
and opto- electronic properties that do not exist in their bulk counterparts, which
renders them attractive for high-performance photodetection applications. The key
parameters of the photoconductors/phototransistors based on TMDs are
summarized in Table 1.
Table 1. Performance of photoconductors/phototransistors based on 2D TMDs.
ML: Multilayer, CVD: Chemical vapor deposition, FL: Few-layer.
Active layer Spectral range Responsivity Response Detectivity Ref.
[AW-1] times [Jones]
[ms]
-3
Mechanically Visible 7.5 ×10 50 – [88]
exfoliated 1L
Mechanically Visible 880 600 – [89]
exfoliated 1L
Mechanically UV-visible- 0.12 – ≈1011 [91]
exfoliated ML NIR
CVD-grown 1L (inVisible 2.2 ×103 2.2 ×105 – [95]
vacuum)
-3
CVD-grown 1L Visible 1.1 ×10
> >103 – [98]
CVD-grown FL Visible 0.57 7 × 10-2 ≈1010 [99]
Liquid exfoliated FL UV-visible- 3.6×10-5 60 – [100]
MoS2 NIR
Magnetron sputtered Visible-NIR 1.8 – 5×108 [102]
5L
Solution-synthesized UV-visible 6.3×10-5 20 4.2× 108 [103]
2–4L
Self-limiting grown Visible 0.55 0.2 – [104]
FL
(after laser
micromachining)
Mechanically exfoliated ML Visible 342.6 > >103 – [105]
(local bottom gated)
Me MechanicallyVisible 104 10 7.7×1011 [106]
exfoliated 1–2L (HfO2
encapsulation)
Mechanically Visible-NIR 2570 1.8 2.2×1012 [107]
exfoliated FL
(enhanced by
ferroelectrics)
Mechanically Visible 5000 2 – [108]
exfoliated 10L
(MESFET structure)
APTES doped FL Visible 5.75×103 – 4.47×109 [111]
MoSe2 Mechanically Visible 97.1 15 – [120]
exfoliated FL
CVD-grown 1L Visible 2.15×10-4 25 – [124]
CVD-grown 1L Visible 1.3×10-2 60 – [125]
CVD-grown ML Visible 93.7 400 – [126]
WS2 CVD-grown FL Visible 9.2×10-5 5.3 – [129]
Mechanically exfoliatedVisible 884 > >103 – [130]
ML
(in NH3 atmosphere)
CVD-grown 1L (inVisible 1.88 ×10-2 <103 – [131]
vacuum)
PLD-grown ML (inUV-visible- 0.70 9.9 ×103 2.7 ×109 [132]
vacuum) NIR
Magnetron SputteredUV-visible 53.3 – 1.22 ×1011 [133]
ML
WSe2 OTS-doped FL Visible 1.45× 104 – 5.3 ×1010 [111]
CVD-grown 1L Visible-NIR 1.8×105> 5×103 >1014 [134]
Mechanically Visible 7 1× 10-2 – [135]
exfoliated 3L
CVD-grown 1L Visible 1.1 > >103 – [136]
PPh3-doped FL onVisible 1.27×106 2.8 – [137]
hBN
MoTe2 Mechanically Visible 2560 – – [141]
exfoliated FL
HfS2 Mechanically Visible 890 38 1.3 ×1010 [142]
exfoliated FL
ReSe
M Mechanically
2 exfoliatedVisible 95 68 – [144]
1L
(in O2 environment)
Mechanically Visible 55.5 96 – [145]
exfoliated 1L (Mo-
doped, in NH3
environment)
ReS2 CVD-grown FL Visible 16.14 > >105 – [146]
Mechanically Visible > >103 98 – [147]
exfoliated FL
CVD-grown 2L Visible 604 2 4.44×1010 [148]
4
Mechanically Visible 8.86 ×10
> >105 1.18×1012 [149]
exfoliated FL
PPh3 and APTES co-Visible-NIR 1.18 ×106 30 – [150]
doped FL
Mechanically exfoliatedVisible 2.5 ×107 670 – [151]
ML (treated by O2
plasma)

3.1.1 Molybdenum Disulphide (MoS2)


Single-layer MoS2 possesses striking properties such as a direct bandgap of ≈ 1.8
eV,[2,29] large carrier mobility above 200 cm2 V–1 s–1[7,29,32], a high-current ON/OFF
ratio of up to 1×108,[2] strong photoluminescence (PL),[7,32] good chemical stability and
mechanical flexibility, ease of processing, and so on.[24,28,29,38] These appealing features
make single- or few-layer MoS2 the most extensively studied 2D layered
semiconductor with great potential in photodetection. A phototransistor based on a
mechanically-exfoliated single-layer MoS2 nanosheet with a typical FET configuration
was first reported by Yin et al.[88] (Inset in Figure 2a). As shown in Figure 2a, the
phototransistor exhibits a distinct response with a cut-off wavelength at 670 nm, which
agrees well with the bandgap of the single-layer MoS2 (≈ 1.8 eV). This device achieved
a maximum responsivity of ≈ 7.5 mAW–1 in the electron accumulation region, which
can be further tailored by drain and gate voltages, along with a fast response speed
within 50 ms. Later, Lopez-Sanchez et al. demonstrated a similar phototransistor
made from exfoliated single-layer MoS2 with a greatly enhanced responsivity up to
880 A W–1 in the electron depletion region (Figure 2b) and a response cut-off
wavelength at 680 nm.[89] This significant improvement is explained by the improved
carrier mobility and contact quality, as well as the positioning technique. Importantly,
this device possesses a remarkably low NEP of down to 1.5×10-15 W Hz-1/2, associated
with its reduced dark current at the depletion region, due to the 1.8 eV bandgap that
suppresses the thermally activated carriers and the efficient electrostatic tunability
in the atomically thick layer. However, the response time of the device is slow (on
the order of several seconds), which is influenced dramatically by the surroundings
of MoS2. With the help of a short gate pulse to discharge trapped charge carriers,
the response times can be reduced to less than 0.6 s. In contrast to the first-reported
MoS2 phototransistor,[88] the photocurrent displays a sublinear dependence on light
intensity. This behavior, combined with the tunable response speed influenced by
the surroundings of MoS2, confirms that charge trapping in MoS2 and/or at the
MoS2/SiO2 interface plays a major role in the sensing process.
The layer number of MoS2 has a significant influence on its band structure,
consequently affecting the photodetection performance.[29,30] Lee and co-workers
have fabricated transparent top- gate phototransistors based on exfoliated single-,
double-, and triple-layer MoS2.[90] Their studies show that the optical bandgap for
the single-layer MoS2 is ≈ 1.82 eV, while the values reduce to ≈ 1.65 and ≈ 1.35 eV
for the double- and triple-layer counter- parts, respectively, as shown in Figure 2c.
FEATURE ARTICLE

As a result, triple-layer MoS2 shows good detection performance for red light, while
single- and double-layer MoS2 are more suitable for green light detection. This
bandgap reduction therefore allows multilayer MoS2 to detect light with an
extended wavelength range up to NIR.[91] Although benefitting from higher light
absorption due to increased thickness, the phototransistor based on multilayer MoS2
exhibits an inferior responsivity of ≈ 100 mA W–1, which might be associated with
the direct bandgap of multilayer MoS2.[92,93] Another work by Khan et al. also
demonstrated that the responsivity and response speed are highly dependent on the
layer number of MoS2 as well as the environment.[94]
Two-dimensional MoS2 prepared with different methods usually have distinct
qualities that are crucial to photodetector performance. A phototransistor based on
CVD-grown monolayer MoS2 reported by Zhang et al. [95] showed a maximum
responsivity as high as 2200 A W–1 in vacuum, which was reduced to 780 A W–1 in
air. Taking into account the large surface-to-volume ratio of MoS2, adsorbates found
at the surface of MoS2 and/ or at the MoS2/substrate interface from ambient air
can act as p-type dopants with additional Coulomb potentials, [96,97] causing
increased carrier scattering and decreased electron mobility and responsivity in air.
The adsorbates could also affect the photoresponse by serving as recombination
centers for photocarriers, leading to a decreased responsivity, yet faster photocurrent
relaxation. The defects and charge impurity states inside the bandgap of MoS2,
which act as trap centers for photocarriers, are considered to be responsible for the
observed persistent photoconductance. In another study, Perea-López and
colleagues also built a photodetector based on CVD-grown monolayer MoS2 with
a much lower responsivity of only 1.1 mA W–1.[98] The huge variation in
responsivities presented in the two studies highlight the important roles of contact
resistance and active contact region in high-performance photodetectors. CVD-
grown few-layer MoS2 has also been exploited as a photodetector for use in harsh
environments.[99] The device exhibits an obvious photocurrent/ dark current ratio
of ≈ 10, even at 200 ℃. MoS2 nanosheets prepared from other methods, such as
liquid exfoliation, [100,101] magnetron sputtering,[102] and solution synthesis [103] have
also been employed in photodetectors. However, these devices usually exhibit
inferior responsivities to their counterparts based on mechanically-exfoliated or
CVD-grown MoS2.
Both photoconductive and photogating effects were found to contribute to the
photocurrent in single and bilayer MoS2 photodetectors.[92] The two effects exhibit
largely different response times, which offers the possibility to identify them in
the photocurrent generation processes. The photogating effect is a slow process and
shows an obvious dependence on the gate voltage, which likely arises from long-
lived charge trap states at the interface of the MoS2 and SiO2 surface hydrated by
water molecules. From the dynamics of the photocurrent measurement, the density
of these trap states is estimated to be ≈ 1015 cm-2. In contrast, the photoconductive
effect is a fast process with negligible gate voltage dependence, which is attributed
to both shallow trap states and mid-gap states as a result of structural defects in
MoS2 or disorder. By modulating the incident light illumination at a faster
frequency than the photogating effect, the photoconductive response of the device
can be studied by a lock-in technique. A model based on mid-gap states and hole-
trap states is proposed to fit the obtained experimental data, from which the density
of hole-trap states is determined to be ≈ 5×1010 cm–2.
Recently, several strategies have been developed with the aim of enhancing the
performance of MoS2 photodetectors. [104–111] Lu et al. demonstrated a facile and
effective focused laser thinning method to achieve micropatterning and localized
modification of multilayer MoS2 films.[104] Due to surface oxidation or oxygen
doping introduced, a photodetector based on modified MoS2 shows a promoted
photoresponse with a responsivity increase of several fold. The photoresponse of a
multi- layer MoS2 phototransistor can be greatly enhanced by using a local
bottom-gate structure, as recently reported by Kwon et al. [105] With this unique
design, the device shows significantly improved photocurrent compared to those of
previously reported global-gated multilayer MoS2 phototransistors.[90,91] The
accumulation of photoexcited holes due to the large tunnel barrier imposed by the
gate at ungated channel regions reduces the potential barrier for electrons, resulting
in increased thermionic current in the electron depletion region. On the other hand,
improved photocurrent in the electron accumulation region is associated with a
decreased tunnel barrier for photoexcited holes and suppressed dark current due to
the series resistance in the ungated regions. As a result, the responsivity increases
drastically—by three orders of magnitude-to 342.6 A W-1. Afterwards, Kufer and
co-workers found that encapsulation with atomic-layer-deposited (ALD) HfO2 can
significantly enhance both the electronic and optoelectronic performance of
monolayer or few-layer MoS2 phototransistors.[106] The encapsulated MoS2 exhibits
enhanced n-type doping with vanishing hysteresis behavior in transfer curves.
This is due to reduced surface adsorbates that serve as charge-trapping centers and
additional electrons induced by positively fixed charges inside the top-oxide layer,
as well as improved carrier mobility originating from the removal of extrinsic
charged impurities and the quenching of homopolar phonon modes. The improved
carrier mobility, combined with lowered contact resistance, thus gives rise to a
much-improved responsivity (Figure 2d) and response speed, which can be
effectively tuned by the gate voltage. The responsivity ranges from ≈ 10 to ≈ 104
AW–1 and response times are in the range of ≈ 10 ms to10 s. Meanwhile, Wang
et al. reported performance enhancement in a multilayer MoS2 phototransistor
driven by a ferroelectric P(VDF-TrFE) gate dielectric.[107] The stable remnant
polarization of P(VDF)-TrFE provides an ultrahigh local electrostatic field in the
channel; this ensures MoS2 is fully depleted and leads to an ultralow dark current.
Based on this design, the device demonstrates a significantly increased
responsivity up to 2750 A W-1 and a high detectivity of ≈ 2.2×1012 Jones. The
response speed is also improved with rise and decay times of ≈ 1.8 and ≈ 2 ms,
respectively, due to the encapsulation or passivation of the surface trap states on
MoS2 by the fluorine or hydrogen atoms from polarized P(VDF-TrFE). More
importantly, the authors found that the detection range of the ferroelectric-driven
phototransistor extended from ≈ 900 nm up to ≈ 1550 nm (Figure 2e), which is
associated with the change in the MoS2 bandgap[112,113] and defects in MoS2 flakes,

FEATURE ARTICLE
both induced by the external electrostatic field. Recently, Lee and colleagues
demonstrated that the influence of interface traps or an on-state gate voltage on
carrier transport behavior in few- layer MoS2 can be minimized by employing a
metal-semiconductor field-effect transistor (MESFET) structure, where the NiOX
contact can form a van der Waals Schottky junction with MoS2.[108] This type of
structure enables high carrier mobilities of 500-1200 cm2 V-1 s-1 at low threshold
voltages due to a reduced charge scattering effect. Thanks to the high carrier
mobility, the device reaches a high responsivity of up to 5000 A W-1 in the transistor
ON state and fast response times of 2 ms (Figure 2f). This response time is superior
to those of a MoS2 phototransistor with metal-insulator- semiconductor FET
(MISFET) configuration, which are >250 ms. Electron doping of single- and few-
layer MoS2 by surface functionalization of Cs2Co3 or an amino-
propyltriethoxysilane (APTES) monolayer has also been proven to be an efficient
way to improve the performance of MoS2-based phototransistors.[110,111] Usually,
these sur- face functionalizations lead to an enhancement in photocurrent of several
folds to tenfold, which is attributed to two phenomena. Firstly, electron doping
induces the formation of negative trions, i.e., quasiparticles comprised of two
electrons and one hole, in MoS2 under illumination. This can significantly reduce
the recombination of photo- excited electron–hole pairs or excitons. On the other
hand, surface functionalization can effectively increase carrier mobility in MoS2
due to suppressed impurity scattering, which is desirable for fast photocarrier
transportation. As a consequence, a high responsivity of up to 5.75×103 A W–1 and
a detectivity close to 1010 Jones have been demonstrated in a APTES-doped MoS2
phototransistor.[111]
Photodetectors with broad response bandwidth are highly desired for some
applications. By using the ultrafast two-pulse photovoltage correlation (TPPC)
technique, Wang et al. showed that the intrinsic response times of monolayer
MoS2 can be as short as 3 ps, enabling wide photodetection bandwidths of up to
300 GHz.[114] The fast response speed is explained by the short lifetime of the
photoexcited carriers due to the dominant defect-assisted recombination.
Figure 2. a) Drain current (Ids) of a single-layer MoS2 phototransistor as a function of
excitation wavelength of the illumination source at a constant optical intensity. Inset
shows an optical image of a typical device made of single-layer MoS2. Reproduced with
permission.[88] Copyright 2012, American Chemical Society. b) Responsivity of a MoS2
phototransistor as a function of incident power. Inset shows the 3D schematic view of the
single-layer MoS2 photodetector. Reproduced with permission.[89] Copyright 2013, Nature
Publishing Group. c) Photon energy-dependent 𝛥Qeff plots indicate the approximate optical
energy gaps to be 1.35, 1.65, and 1.82 eV for triple-, double-, and single-layer MoS2
nanosheets, respectively. Reproduced with permission.[90] Copyright 2012, American
Chemical Society. d) Power-dependent responsivity before and after HfO2 encapsula- tion
of a bilayer MoS2 phototransistor. Reproduced with permission.[106] Copyright 2015,
American Chemical Society. e) Responsivity of a P(VDF-TrFE) polarization-gating triple-
layer MoS2 photodetector as a function of light wavelength from 500 nm to 1550 nm at
Vsd=1 V and P=100 μW. Inset shows the 3D schematic view of the polarization-gating
triple-layer MoS2 photodetector. Reproduced with permission.[107] Copyright 2015, Wiley-
VCH. f) Dynamic photoswitching behavior of a MoS2 phototransistor with MESFET
configuration. Reproduced with permission.[108] Copyright 2015, American Chemical
Society.
Photo-Thermoelectric Effect in MoS2. In addition to the photoconductive effect and
photogating effect, the photo-thermoelectric effect has been found to contribute to
the photoresponse process in a single-layer MoS2 FET.[115] In this study, the authors
examined the photoresponse of the FET with no external bias via scanning
photocurrent microscopy (SPCM) measurements and found a strong and tunable
photo-thermoelectric effect. An AFM image of the studied device is shown in
Figure 3a. During the test, electrode 3 was connected to a current-to-voltage
amplifier and the other electrodes were grounded. Under above- bandgap
illumination, remarkable photocurrent was observed, even when the laser spot was
focused inside the area of the electrodes (Figure 3b). To further clarify the origin
of the photocurrent, they performed measurements under sub-bandgap illumination
and demonstrated qualitatively similar photoresponse behavior with lower
photocurrent values. These results clearly indicate that the photoresponse process
is dominated by the photo-thermoelectric effect. The Seebeck coefficient S was
determined by dividing the measured photo-thermoelectric voltage VPTE by the
temperature gradient, which can be estimated via a finite elements analysis
calculation. Figure 3c shows the calculated S as a function of the gate voltage. Note
that a negative value of S is expected for an n-type semiconductor. Evidently, the S
of the single-layer MoS2 reaches its maximum ≈3×105 μV K–1 in the electron
depletion region and rapidly decreases by two to three orders of magnitude in the
electron accumulation region, which is significantly larger than that of graphene (±4
to 100 μV K–1).[115] The large and tunable S makes single-layer MoS2 a promising
material for applications such as on-chip thermopower generation waste thermal
energy harvesting. In another work, Zhang et al. also found that the hot-carrier-
assisted-photo-thermoelectric effect dominates the photocurrent generation in the
electron depletion region in a multilayer MoS2 transistor with external bias, while
the photovoltaic effect at the MoS2-metal interface plays a major role in the electron
depletion region.[116]

Figure 3. a) AFM image of phototransistors based on MoS2. b) SPCM map. The


photocurrent (linear colorscale) is collected simultaneously with the intensity of
the reflected light, which is used to detect the contour of the electrodes (solid black
lines). c) Estimated Seebeck coefficient S versus gate voltage. The gray shaded
area is the uncertainty in S due to the uncertainty in the estimation of the
temperature gradient. (a–c) Reproduced with permission.[115] Copyright 2013,
American Chemical Society.
3.1.2. Other TMDs
In addition to MoS2, many other 2D layered TMDs have been explored for
photodetection, including MoSe2, WS2, WSe2, MoTe2, HfS2, ReS2 and ReSe2. In
this section, we introduce photoconductors/phototransistors based on these 2D
layered TMDs.
Single-layer MoSe2 shows many appealing properties similar to those of single-
layer MoS2, such as a direct bandgap of ≈1.5 eV,[117] strong PL,[54] and a large
exciton binding energy.[118] Rapid advances in preparing high-quality single- and
few-layer MoSe2 by mechanical exfoliation[119,120] or CVD techniques[121–123] have
expanded the range of photodetection applications. For example, phototransistors
based on CVD- grown monolayer MoSe2 were reported by Chang et al.[124] and
Xia et al.[125] with responsivities on the order of mA W–1, several orders of
magnitude lower than that of a device based on CVD-grown monolayer MoS2.[95]
However, fewer defects or charge impurity states within MoSe2 compared to
MoS2 give rise to much faster response times on the order of tens of milliseconds.
The responsivity can be drastically increased to ≈93.7 A W–1 at the cost of reduced
response speed by using CVD-grown multilayer MoSe2 as a photoactive material,
due to a large density of trap states within the bandgap of the MoSe2.[126] Similarly,
Abderrahmane et al. demonstrated a phototransistor made from exfoliated few-
layer MoSe2 with a thickness of ≈20 nm.[120] Their device reaches a maximum
responsivity of up to 97.1 A W–1 near the threshold gate voltage and response times
of tens of milliseconds.
Single- and few-layer WS2 have been synthesized via various methods and
employed in the field of photodetection.[127,128] Perea-López et al. have studied the
photoresponse of CVD- grown few-layer WS2, which is highly dependent on the
photon energy from light excitation.[129] The device shows a low responsivity of ≈ 92
μA W-1 and a fast response speed of ≈5 ms. Afterwards, Huo and co-workers found
that the photoresponse of exfoliated multilayer WS2 is strongly dependent on the
sur- rounding gaseous environment (Figure 4a).[130] The responsivity increases
from tens of A W–1 in vacuum to a maximum of 884 A W–1 in an NH3
atmosphere, which can be ascribed to the charge transfer between the physically
absorbed gas molecules and the WS2. This charge transfer modulates the WS2
doping level and prolongs the lifetime of photocarriers, leading to an enhanced
responsivity. A similar phenomenon in CVD-grown monolayer WS2 was recently
reported by Lan et al.[131] In this study, the device had a responsivity of 18.8 m A
W–1 in vacuum, which is drastically reduced in air to 0.2 μA W–1. Meanwhile,
response times are on the order of subseconds in vacuum and less than 4.5 ms in
air. More recently, centimeter-sized multi- layer WS2 films tens of nanometers in
thickness have been successfully prepared by pulsed-laser deposition[132] or
magnetron sputtering techniques,[133] respectively. Photodetectors based on these
films show decent responsivities between 0.51 A W–1 and 53.3 A W–1 with
broadband photoresponses. Considering the facile and scalable fabrication
processes, as well as the large- size, homogeneous, and transferable layered films
obtained on various substrates, these preparation methods may open up new
opportunities towards practical applications in the future.
Single- and few-layer WSe2 have also demonstrated their potential in
photodetection applications. Zhang et al. have studied the effect of metal contacts
with different work functions on the photoresponse of CVD-grown monolayer
WSe2.[134] The device with Pd contacts reaches a maximum responsivity of up to
≈1.8×105 A W–1, while the value decreases by several orders of magnitude for the
device with Ti contacts in air. Interestingly, the response times for the device with
Ti contacts are less than 23 ms, much faster than the Pt-contacted device (tens of
seconds), as shown in Figure 4b. The significant variation in device performance is
attributed to the large difference in Schottky barriers between the different metals
and WSe2, indicating the critical role of metal contacts in highly sensitive
photodetectors. In another work, Pradhan and colleagues demonstrated a
photodetector based on mechanically exfoliated tri-layer WSe2 with responsivities
of 7 A W–1 and fast response speed on the order of 10 μs.[135] Very recently, a WSe2
photodetector with graphene electrodes that achieves a decent responsivity of ≈1.1
A W–1 was reported, suggesting good electrical contact between graphene and
WSe2.[136] Similarly, doping of few-layer WSe2 can significantly affect its
optoelectronic performance.[111,137] Specifically, Jo et al. demonstrated a high-
performance WSe2-based phototransistor, achieved by converting the conduction
type from the p- to n-channel by triphenylphosphine (PPh3) n-doping.[137] The
responsivity of a device with Ti contacts increased by 160-fold, from 2.70×103 to
4.31×105 A W–1, at a low light level upon doping. Further performance
enhancement was attained by insertion of an h-BN layer underneath the WSe2
channel, providing a charge-free environment that can greatly enhance the carrier
mobility. Eventually, the device showed a responsivity as high as 1.27×106 A W–1
and fast response times of 2.8 ms.

Figure 4. a) The responsivity R of a multilayer WS2 phototransistor as a function of


light intensity under various gas atmospheres. Reproduced with permission.[130]
Copyright 2014, Nature Publishing Group. b) Photoswitching behaviors of the Pd- and
Ti-contacted WSe2 photodetectors. Reproduced with permission.[134] Copyright
2014, American Chemical Society. c) 3D schematic view of an ReS2
photodetection device. The light is illuminated on the ReS2 channel, and the light
polarization is controlled by the half wave plate. d) The photocurrent with drain
bias of 1 V (red square) and absorption (green circle) measured under different
polarization angle of green light and plotted in polar coordinates. The blue lines are
the fitting results using sinusoidal function. The consistency between photocurrent
and absorption indicates the intrinsic linear dichromic response of ReS2. (c,d)
Reproduced with permission.[147] Copyright 2016, Wiley-VCH.
Due to its outstanding optical and electrical properties, ultrathin MoTe2, a recently
FEATURE ARTICLE

FEATURE ARTICLE
introduced 2D material, has received much attention.[138–140] Yin et al. studied the
influence of metal contacts on the electrical properties of a phototransistor based
on mechanically exfoliated few-layer MoTe2.[141] Under optimum conditions, the
device reached a responsivity of up to 2560 A W–1.
More recently, Xu and co-workers successfully isolated single- and few-layer HfS2
flakes via mechanical exfoliation.[142] Experiments and theoretical simulations
revealed that ultrathin HfS2 is an indirect gap material with a bandgap of ≈2 eV.
Moreover, they found that the electrical properties of the HfS2-based
phototransistors are greatly influenced by flake thickness and metal contacts. With
optimized flake thickness, the responsivity reaches a maximum of ≈890 A W–1 in
the transistor ON state with Au contacts and decreases tens of times with Ti
contacts. Interestingly, the response times show strong back-gate dependence with
distinct trends between the devices with different contacts. The different
photoresponse behaviors are attributed to different Schottky barriers between the
flake and the contacts, which are crucial to the separation and transportation of
photocarriers.
Different from most layered TMDs with relatively high crystal symmetry, Re
dichalcogenides, such as ReS2 and ReSe2, are anisotropic semiconductors
crystallized in a distorted 1T in-plane structure.[143] Their extremely anisotropic
electrical, optical, and mechanical properties render these compounds quite
interesting for novel electronics and optoelectronics. Yang et al. found that the
bandgap and carrier mobility of exfoliated ReSe2 are highly modulated by their
layer thickness, allowing the opportunity to engineer the electronic and
optoelectronic properties of ReSe2-based devices.[144] A phototransistor based on
single-layer ReSe2 shows a remarkable photoresponse with a responsivity reaching
95 A W–1 and fast response times on the order of tens of milliseconds. Similar to
MoS2 and WS2 photodetectors,[94,95,130,131] the optoelectronic properties of the
device are significantly influenced by the surrounding environment, which has also
been found in a photodetector based on few-layer-Mo-doped ReSe2 in another
work.[145] This phenomenon is attributed to the charge transfer between ReSe2 and
gas molecules; this affects the doping level in ReSe2 as well as the lifetimes of the
photocarriers. A big disadvantage for the application of ReSe2 in photodetectors is
its poor air stability, which degrades its photodetecting performance. Therefore,
passivation or encapsulation schemes for real-world applications are in high
demand.
One particular case is illustrated by ReS2, which maintains a direct bandgap of 1.5
eV irrespective of layer thickness. Both CVD-grown and exfoliated few-layer ReS2
have been extensively explored as photodetectors, exhibiting responsivities ranging
from 16.14 A W–1 to 8.86×105 A W–1, and response times between several
milliseconds to hundreds of seconds.[146–149] Such a huge variation in device
performance can be attributed to three possible reasons: (i) different light
absorption due to varied flake thicknesses, (ii) disparate gain enhancement
depending on the density of trap states in the flakes prepared by different methods,
and (iii) diverse device architectures as well as different contact qualities. An
obvious weakness of these devices is that the current normally cannot be recovered
to their dark current levels after removing the illumination, due to a slow
recombination rate of photocarriers. This issue can be addressed by the application
of a short-gate voltage pulse, which actively purges the charge carriers to reset the
device.[77] Specifically, the ReS2 photodetector can be used to detect polarized light
with high sensitivity, benefiting from the crystal structure anisotropy (Figure 4c and
d).[147] Through electron doping by PPh3 together with APTES, the responsivity of
a ReS2-based phototransistor can be drastically enhanced by several orders of
magnitude, from 3.97×103 to 1.18×106 A W–1.[150] The device also demonstrates
broadband photodetection and fast response times on the order of tens of
milliseconds. This improvement is explained by increased carrier mobility due to
suppressed interfacial carrier scattering, in addition to reduced contact resistance at
the metal-ReS2 junction as a result of the n-doping effect and blocked charge puddle
effect.
More recently, Shim et al. found that O2 plasma treatment can significantly improve
both the electronic and optoelectronic properties of a few-layer ReS2 based
transistor.[151] The maximum responsivity attained in this work is as high as 2.5
×107 A W-1, among the highest values achieved by 2D layered semiconductors with
back-gated transistor geometry. The high responsivity is attributed to high
absorption by the flakes due to their direct bandgap and relatively large thickness
(≈30 nm). The response times are on the order of several to tens of seconds and
decrease with prolonged plasma treatment duration, due to the trap states formed
in the bandgap of ReS2 at the surface region by the treatment. These trap states
increase the photocarrier recombination rate and consequently reduce their
lifetimes, leading to reduced response times.
In summary, photoconductors/phototransistors based on single- or few-layer TMDs
show widely varied detection performances with responsivities ranging from ≈10-7
A W–1 to ≈107 A W–1 and response times between 10–5 s to 103 s. Higher
responsivities are obtained in devices where trap states exist in TMDs and/or at
TMD/SiO2 interfaces at the cost of slower response speeds. Moreover,
performance is highly influenced by the TMD’s layer number, preparation method,
electrode contact qualities, and surrounding environment. Further improvement in
device performance is achievable via various techniques, including the introduction
of local- gated, ferroelectric-polarization-gated, or MESFET device geometries,
encapsulation with an HfO2 layer, or doping techniques
3.2. Group IIIA, IVA, IVB and Ternary Metal Chalcogenides
In addition to the TMDs discussed, other metal chalcogenides have been explored
in the field of photodetection, including group IIIA metal (Ga, In) chalcogenides,
group IVA metal (Ge, Sn) chalcogenides, group IVB metal (Ti, Zr, Ha)
trichalcogenides, and ternary metal chalcogenides. In this section,
photoconductors/phototransistors based on these metal chalcogenides are
discussed and device performance is summarized in Table 2.
Table 2. Performance of photoconductors/phototransistors based on other 2D metal
chalcogenides and few-layer BP. FL: Few-layer, ML: Multilayer, LN: liquid nitrogen,
CVD: Chemical vapor deposition, NRs: Nanoribbons.

3.2.1. Group IIIA (Ga, In) Metal Chalcogenides


Similar to TMDs, group IIIA metal chalcogenides are layered materials with weak
van der Waals interactions between neighboring layers. In Ga chalcogenides with
1:1 stoichiometry such as GaS, the atoms are arranged following the repeating unit
S–Ga–Ga–S in the in-plane direction, which is different from single-layer TMDs.
In the out-of-plane direction, the atoms are arranged in a graphene-like honeycomb
lattice.[152] The unique crystal structure leads to distinct band structures of the Ga
chalcogenides, making them ideal candidates for photodetection.
Photodetectors based on exfoliated multilayer Ga chalcogenides, including
GaSe,[153] GaS,[152] and GaTe,[154,155] have recently been realized on rigid or flexible
substrates (Figure 5a). The GaSe-based device shows a responsivity of 2.8 A W–1
and a response speed of 0.02 s,[153] while the GaS-based devices made on SiO2/Si
and polyethylene terephthalate (PET) substrates show responsivities of up to 4.2 A
W–1 and 19.2 A W–1 in the UV range, respectively (Figure 5b).[152] The detectivities
achieved by the GaS-based devices are in the range of ≈1013-1014 Jones (Figure 5b),
surpassing commercial Si or InGaAs devices. Interestingly, the GaTe-based devices
exhibit significantly increased responsivities between 274.3 A W–1 and 104 A W–1
and response speeds down to several milliseconds.[154,155] The much larger
responsivities can be explained by the higher absorption coefficient and more
efficient photoexcited electron–hole pair generation thanks to the direct bandgap of
GaTe.[154,155] The photoresponse displays a sublinear dependence on the excitation
intensity, indicating that the trap states presented in GaTe and/or at GaTe/substrate
interfaces play a crucial role in the photocurrent response. Similar to
TMDs,[94,95,130,131,144,145] the electronic/optoelectronic properties of 2D Ga
chalcogenides are also sensitive to surface adsorbates such as moisture, oxygen
molecules, and other gaseous molecules.[156] This offers the opportunity to
improve the photoresponse in terms of responsivity, EQE, and response speed by
employing the appropriate operating conditions. Moreover, Ga vacancies play a
critical role in the electronic/optoelectronic properties of a phototransistor based on
exfoliated GaTe.[157] By suppressing the thermally activated Ga vacancies at liquid
nitrogen temperature in vacuum, the device exhibits an improved responsivity of as
high as 800 A W–1 and a decreased response time. Recently, Cao et al. demonstrated
significant enhancement in responsivity (up to 5000 A W–1) by shrinking the
electrode spacing distance to 290 nm in few-layer GaSe based photodetectors.[158]
Their device shows response times on the order of tens of milliseconds, which can
be further reduced to hundreds of micro-seconds by employing graphene
electrodes. In addition, photo- detectors with regular responsivities between 0.017
A W–1 to 1.7 A W–1 have also been reported based on single- or few-layer Ga
chalcogenides synthesized from other routes, including vapor phase mass transport
(VMT),[159] PLD,[160] vapor phase deposition,[161] and CVD[162] methods.
Apart from Ga chalcogenides, In chalcogenides, including In2Se3[163–167] and
InSe,[168–172] have also shown great promise in photodetection applications. The
In2Se3 compounds have multiple structural phases (i.e., α, β, γ, δ, and κ) that affect
their electronic and optoelectronic properties.[173] α-Phase In2Se3, one of the two
common forms, has a direct bandgap of about 1.3 eV,[174,175] making it suitable for
visible light harvesting in optoelectronics. Single- or few-layer α-In2Se3 flakes have
been successfully synthesized on various substrates by van der Waals epitaxy.[163]
A photodetector based on such a flake shows an efficient photoresponse with a
responsivity of ≈2.5 A W–1. Afterwards, Jacobs-Gedrim et al. demonstrated
extraordinary photoresponse behavior from UV-visible to NIR regions in exfoliated
α-In2Se3 with ≈10 nm thickness.[164] Their device has a responsivity as high as 3.95
×102 A W–1, fast response times on the order of tens of milliseconds, and a high

FEATURE ARTICLE
specific detectivity of over 1012 Jones. The excellent photoresponse is believed to
be attributed to the direct bandgap of In2Se3 and its large surface area-to-volume
ratio, while the reduced carrier recombination as a result of the self-
terminated/native-oxide-free surface leads to fast response. Zhou and colleagues
have grown high- quality monolayer α-In2Se3 flakes via a physical vapor deposition
(PVD) method and examined their optoelectronic properties.[165] A photodetector
made from this flake demonstrates a similar photoresponse with the devices based
on exfoliated ones, with a responsivity of 340 A W–1 and response times of ≈10 ms.
Very recently, Island et al. have found that the photocurrent generation mechanism
is a combination of the photoconductive effect and the photogating effect in isolated
multilayer α-In2Se3.[166] The photoconduction is gate-independent, while
photocurrent from the photogating effect can be strongly modulated with the back
gate. Specifically, the dominant mechanism can be effectively tuned from fast
conduction in the transistor OFF state to high gain photogating in the ON state via
the back gate. The strong photogating effect is attributed to the long-lived states
from intrinsic defects and native oxide at the In2Se3 flake surface, which trap
photogenerated holes and effectively gate the flake with increased photocurrent.
This strong photogating effect, along with the high absorption coefficient due to
the direct bandgap of In2Se3, give rise to an extremely high responsivity of ≈105 A
W–1 and an inferred detectivity of ~3.3×1013 Jones in the In2Se3 phototransistor.
Recently, Zheng et al. developed a method combining microintaglio printing with
van der Waals epitaxy to efficiently pattern various 2D metal chalcogenides,
including In2Se3 crystal arrays on mica substrates.[167] Precise control of the
crystallization, thickness, position, orientation, and layout of patterned 2D crystals
has been accomplished. In addition, the patterned In2Se3 crystal arrays can be
integrated and packaged into flexible photodetectors with a prominent responsivity
of ≈1650 A W-1. This study has opened a new opportunity for batch fabrication and
integration of photo-detectors, suggesting great promise for next-generation
flexible, wearable, and integrated electronics and optoelectronics.
Different from other group IIIA metal chalcogenides, InSe has a direct bandgap of
≈1.3 eV in bulk form, while a direct-to-indirect transition occurs when reducing its
thickness to a few nanometers.[176] Lei at al. demonstrated efficient photocurrent in
exfoliated multilayer InSe with a varied number of layers under photoexcitation in
the range of 400–800 nm.[168] The tail of the photocurrent vs wavelength spectra for
a 10-layer InSe photodetector can be fit well with a parabola, suggesting the indirect
nature of the bandgap in few-layer InSe. The device shows a responsivity of 34.7
mA W–1 and a fast response time of 488 μs. Subsequently, Tamalampudi and co-
workers have fabricated phototransistors based on exfoliated few-layer InSe on
both rigid SiO2/Si and bendable PET substrates.[169] Their devices show broadband
photoresponses from the visible to NIR region with responsivities of ≈10 A W–1
(Figure 5c), which can be improved significantly to 157 A W–1 with a back-gate
voltage. The photocurrent increases sublinearly with the incident excitation (Figure
5c) while the response time is increased to several seconds. This indicates that long-
lived trap states caused by the defects and/or charged impurities in InSe and the
adsorbed molecules at the InSe/SiO2 interface play an important role in the
photoresponse. Recently, Feng et al. discovered that multiple reflection
interference at the interfaces in multilayer InSe phototransistors leads to a channel-
thickness-dependent photoresponse based on both theoretical simulations and
experimental results.[170] With an optimal InSe thickness of ≈30 nm, they
demonstrated a high responsivity of up to ≈5.6×104 A W–1 and a detectivity above
1013 Jones in spectra ranging from UV-visible to NIR. The device also showed fast
response times on the order of several milliseconds. Specifically, the response times
of few-layer-InSe-based photodetectors can be dramatically reduced from several
milliseconds to 120 μs by using graphene electrodes, as reported by Luo and
colleagues.[171] The much-improved response speed is attributed to the smaller
Schottky barrier between InSe and graphene as well as the high mobility of
graphene, which facilitates fast transportation of photocarriers. They also
demonstrated effective modulation of the photocurrent and response times via the
back-gate voltage because the work function of the two materials can be easily
regulated by the gate voltage, leading to a tunable Schottky barrier. More recently,
the avalanche effect, which can induce carrier multiplication to greatly enhance the
photoresponse, has been realized in few-layer InSe. This effect was triggered by
exploiting the large Schottky barrier between the InSe and Al contacts to enable the
application of a large bias voltage.[172] Thanks to carrier multiplication, the device
reaches an EQE of 334% with an avalanche gain of 47 at a moderate voltage range
before reverse bias breakdown, while the values increase to 1110% and 152,
respectively, at the voltage range where reverse bias breakdown occurs. The large
Schottky barrier also reduces the dark current to picoamp scale, enabling a large
single to noise ratio (SNR) on the order of tens of decibels. The photodetector
shows a fast response time of 60 μs, about two orders of magnitude faster than a
back-gated InSe photodetector with a similar photoresponse level. Further
integration of plasmonic nanoantennas based on an Al disk can greatly enhance the
photoresponse of the device to improve the EQE by approximately 8 times. The
authors also found that the photocurrent increases significantly in the wavelength
range of 650–750 nm beyond the scattering peak of Al nanodisks, which is
associated with the electron emission from the Fermi level of the Al into InSe. Three
different processes can account for this improvement: For one, the plasmonic
nanodisks can provide a local electromagnetic field enhancement that contributes
to carrier generation in InSe. On the other hand, the decay of the surface plasmons
of the nanodisks produces hot electrons, and those with sufficient energy to
overcome the Al–InSe Schottky barrier can provide additional photocarriers. Third,
electrons emitted over the Schottky barrier can be accelerated to a high kinetic
energy, further contributing to the avalanche process.
3.2.2. Group IVA (Ge, Sn) Metal Chalcogenides
In addition to the TMDs and group IIIA metal chalcogenides discussed, there is a
growing interest in photodetection based on ultrathin group IVA metal
[177–182] [183–185] [186] [187]
chalcogenides, including SnS2, SnSe2, GeSe, and GeS, for
their unique electronic structures and physical properties. In addition, this group
of semiconductors is made up of earth-abundant and environmental-friendly
FEATURE ARTICLE

elements with prominent chemical stability, making them particularly attractive for
practical applications.
Tin dichalcogenides SnS2 and SnSe2 are isostructural with the hexagonal cadmium
iodide (CdI2)-type structure and exhibit indirect bandgaps of ~2.2 eV and ~1.0 eV
in bulk, respectively.[188,189] Both semiconductors have recently attracted much
attention in the field of photodetection. For example, single- and few-layer SnS2
flakes have been successfully exfoliated and employed in transistors and
photodetectors.[177] Experiments and theoretical calculations show that the indirect
bandgap nature remains in SnS2 even with thickness down to a monolayer. A device
with a flake of ≈10 nm thickness reaches a high responsivity of ≈100 A W–1. Similar
to MoS2-based phototransistors,[92] the photoresponse contains two components, i.e.,
a slow component with response times on the order of several to tens of seconds
and a fast component with response times on the scale of tens of milliseconds. The
slow and fast components are likely due to a combination of long-lived extrinsic
traps, such as adsorbates at the SnS2 surface and SnS2/SiO2 interface, as well as
defect states in SnS2. Photodetectors based on ultrathin SnS2 prepared by other
methods, including CVD,[178–180] atmospheric pressure vapor deposition
(APVD),[181] and self-assembled syn- thesis[182] have also been realized on rigid or
flexible substrates and reach significantly varied responsivities from ≈0.2 mA W–1
to ≈100 A W–1. Specifically, ultrathin SnS2 crystal arrays can be grown by a CVD
method via seed engineering, and photodetectors composed of these crystals have
fast response times of 5 μs, as shown in Figure 5d.[178] Importantly, the
performance of phototransistors is highly sensitive to the density of sulfur
vacancies in SnS2. Through temperature-dependent electrical transport
measurements together with theoretical calculations, Huang et al. have verified that
sulfur vacancies generated in the growth process can induce defect levels near the
bottom of the conduction band of SnS2.[179] Accordingly, the photoresponse of a
SnS2-based phototransistor can be significantly improved by diminishing the sulfur
vacancies, as reported by Yang and co-workers recently.[180] Through sulfur
annealing treatment, their device achieves an 30-fold improvement in the
photoswitching ratio, with dark current suppressed by several orders of magnitude
and response times reduced by more than 3 orders of magnitude to less than 1 ms.
Zhou et al. have recently synthesized single-crystalline SnSe2 flakes with
thicknesses down to ≈1.5 nm via an improved CVD method.[183] Their experiments
revealed that few-layer SnSe2 with a thickness of ≈10 nm exhibits an indirect
bandgap of ≈1.73 eV. A photodetector composed of the flake reaches a high
responsivity of ≈1.1×103 A W–1 and fast response times of ≈10 ms. Another work
by Huang et al. demonstrated the synthesis of SnSe2 nanosheets with diverse shapes
and thicknesses varying from ≈10 nm to ≈150 nm.[184] Photoresponse
measurements on a ≈20 nm-thick nanosheet showed a responsivity of ≈1.9 A W-
1
and a current on/off ratio of ≈100. More recently, Yu and colleagues successfully
exfoliated ultrathin 1T-phase SnSe2 flakes with various layer numbers from bulk
single crystals and probed their optoelectronic properties.[185] Their device gave an
evident photoresponse with a decent responsivity of 0.5 A W–1 and a fast response
time of down to ≈2 ms.
Group IVA chalcogenides with 1:1 stoichiometry, such as GeS, GeSe, SnS, and
SnSe, are important p-type semiconductors with a layered crystal structure and an
atom arrangement resembling the distorted NaCl structure.[190] These compounds
have both indirect and direct bandgaps in the range of 0.90–1.65 eV, rendering them
promising absorbers in the visible to NIR spectral region.[191,192] Single-crystal GeSe
nanosheets with thicknesses of ≈300 nm have been successfully synthesized via a
solution-phase method and employed in photodetectors.[186] The device in this work
shows a remarkable photo- response with an obvious anisotropic on/off switching
ratio in directions parallel and perpendicular to the layer, attributed to the large
bonding anisotropy of the layered crystal structure. Recently, Ulaganathan et al.
have isolated multi-layered GeS flakes having ~28 nm thickness and applied them
as channel materials in phototransistors.[187] Their device reached a responsivity of
≈206 A W–1; this value was further increased to as high as ≈655 A W–1 when the
transistor was operated in the ON state.
3.2.3. Group IVB (Ti, Zr, Hf) Metal Trichalcogenides
In this section, we review photodetectors based on ultrathin group IVB metal
trichalcogenides, such as TiS3, ZrS3, HfS3, ZrSe3, and HfSe3. These trichalcogenide
compounds possess direct or indirect bandgaps between ~1.0 eV to ~3.1 eV,
covering the UV to NIR region.[193–195] Island et al. have successfully isolated few-
layer TiS3 nanoribbons (NRs) with thicknesses down to 10–30 nm via mechanical
exfoliation during the transfer process of NRs and probed their photoresponses.[196]
Their device showed a broad photoresponse with a detection range up to 940 nm
(Figure 5e), a high responsivity of 2910 A W–1, and fast response times on the order
of several milliseconds. The photocurrent increased sub- linearly with respect to
incident power; together with decreased responsivity with increasing power, this
suggests that long-lived trap states played a crucial role in the photoresponse.
In another work, Tao and co-workers transferred a network of ZrS3 NRs onto a
flexible substrate such as polypropylene or paper and used it in a photodetector.[197]
The ZrS3 NRs with thicknesses of ≈70 nm were synthesized via a chemical vapor
transport (CVT) method. The fabricated devices exhibited a remarkable
photoresponse in the range of visible to NIR and response times of tens of seconds.
Afterwards, Tao et al. examined the photoresponse of phototransistors based on
single or several ZrS3 NRs.[198] Both devices exhibited responsivities on the order
of tens of A W–1 and response times below 0.4 s, which were superior to those
attained from the device based on the network of ZrS3 NRs.[197] In addition, HfS3,
ZrSe3, and HfSe3 NRs have also been synthesized by CVT methods, as reported
by Xiong and colleagues.[199,200] Photodetectors made from these single NRs
demonstrated impressive photoresponses, with responsivities between 12 mA W–
1
and 530 mA W–1 and response times of less than 0.4 s.
3.2.4. Ternary Metal Chalcogenides
Recently, ternary 2D semiconductors, which offer multiple degrees of freedom for
controlling their physical properties via stoichiometric alteration, have stimulated

FEATURE ARTICLE
much research interest in many fields.[201,202] This emerging class of new 2D
materials has also been employed in photodetection applications.[203–205] For
example, ternary copper indium selenide (CuInSe) with a Cu:In ratio between 1:5
to 1:7 (i.e., γ-phase) is a layered compound.[206] Lei et al. reported successful
exfoliation of few-layer CuIn7Se11 flakes and employed them in photodetection and
photovoltaic applications.[207] Their photodetector with metal-semiconductor-metal
(MSM) configuration gave a strong photoresponse throughout the visible range
spectrum, with a responsivity of 380 mA W–1. Fitting the tail of the photo- response
vs incident photon energy spectra indicated that the CuIn7Se11 with 3-4 layers had
an indirect bandgap of ~1.1 eV, while the value rose to ≈1.4 eV for those with 1-2
layers. Due to the Schottky barrier between the CuIn7Se11 and metal contacts, the
device exhibited a low dark current of ≈1 pA, giving rise to a noise level down to
≈1.3×10–29 A2 and a high SNR of 95 dB.
In another work, Perumal and co-workers fabricated photo- transistors on both rigid
and flexible substrates based on exfoliated few-layered Sn(SxSe1–x)2 flakes with
thicknesses of ≈6 nm and examined their photoresponses.[208] The device on the
rigid substrate showed a prominent photoresponse with a high responsivity of up to
6000 A W–1 (Figure 5f), fast response times of ≈9 ms, and a detectivity exceeding
8×1012 Jones. Notably, their flexible device demonstrated good optoelectronic
properties and excellent bending stability. These studies emphasize ternary 2D
layered semiconductors as highly promising platforms for novel photodetection
applications.
In summary, photodetectors based on these metal chalcogenides show
responsivities (0.2 mA W–1 to 5.6×104 A W–1) and response times (5 μs to tens of
seconds) comparable to those exhibited by TMD-based devices. Moreover, the
wide range of bandgaps available from these metal chalcogenides enables their
operation in a broad spectrum ranging from the UV-visible to NIR region. Element
vacancies can significantly affect the photoresponses of detectors. Further
improvement of the response speed is achievable by using graphene electrodes to
reduce Schottky barriers at the contacts and facilitate fast photocarrier
transportation. In addition, ternary metal chalcogenides can afford multiple degrees
of freedom to regulate their optoelectronic properties, suggesting new platforms for
novel photodetection applications.
Figure 5. a) GaS nanosheet photodetectors on a flexible PET substrate. b) Wavelength-
dependent responsivity and detectivity of the flexible GaS photodetector under 0.5 mW/cm2
irradiance at a bias voltage of 2 V.) a,b) Reproduced with permission.[152] Copyright 2013,
American Chemical Society. Photocurrent (green solid dots) and responsivity (blue solid
squares) of an InSe phototransistor as a function of illumination intensity at Vds = 10 V and
Vg = 0 V. The power laws of R≈P–0.56 and Iph ≈ P0.41 were determined from fitting the
measured data. Reproduced with permission.[169] Copyright 2014, American Chemical
Society. d) Drain-source current IDS of an SnS2 photodetector recorded as a function of time
with a sampling rate of 660 kHz when the laser beam was modulated periodically at a
frequency of ≈800 Hz via an acousto-optic modulator (with the laser beam only partially
deflected). Reproduced with permission.[178] Copyright 2015, American Chemical Society.
e) Photocurrent of a TiS3 NR photodetector measured as a function of different laser
wavelengths. By linearly extrapolating (dashed line) the photocurrent vs. excitation
wavelength data, the bandgap energy is estimated to be 1.2 eV (1010 nm). Inset shows an
AFM image of the device with electrodes spaced by 500 nm. The width of the TiS3 NR is
195 nm. Reproduced with permission.[196] Copyright 2014, Wiley-VCH. f) 3D view of
responsivity mapping of a few-layered Sn(SxSe1–x)2 phototransistor. Reproduced with
permission.[208] Copyright 2016, Wiley-VCH.
3.3. Few-layer Black Phosphorus
Black phosphorus (BP), a stable layered allotrope of phosphorous, is an elemental
p-type semiconductor with a direct bandgap of 0.35 eV and mobilities on the order
of 10 000 cm2 V–1 s–1 in bulk form, and has been studied from the 1980s.[209] At
ambient conditions, orthorhombic BP forms puckered layers held together by weak
interlayer forces with significant van der Waals character[210,211] and can be easily
isolated into thin layers. Unlike its bulk form, few-layer BP shows ambipolar
transport behavior with mobilities up to 1000 cm2 V–1 s–1.[25] Meanwhile, due to the
reduced symmetry in the puckered-layer structure, few-layer BP has been found to
possess remarkable in-plane anisotropic electronic, optical, and phonon
properties.[25,212–215] Previous studies have predicted that the direct bandgap of
few-layer BP depends strongly on the number of layers and the in-layer strain,
where the value increases to over 1 eV for a single layer.[33,34] Thus, the narrow
bandgap renders few-layer BP an appealing candidate for NIR photodetection
(Table 2), complementing the detection gap between zero-bandgap graphene and
large-bandgap TMDs.
Buscema et al. studied the performance of detectors composed of few-layer BP
having 3–8 nm thickness as a function of excitation wavelength, power, and
frequency.[216] The device gave a remarkable photoresponse to excitation
wavelengths from the visible region up to 940 nm (Figure 6a) and demonstrated a
responsivity of 4.8 mA W–1 and fast response times of several milliseconds. Both
the photovoltaic effect and photo thermoelectric effect have been found to
contribute to photocurrent generation in ultrathin BP photodetectors.[217] In this
study, the authors performed scanning photocurrent measurements on a transistor
based on an 8 nm thick, few-layer BP as a function of polarization and gate
voltages. They found that photogenerated electron–hole pairs were separated by a
built- in electric field at Schottky barriers induced by the Fermi level alignment in
the transistor OFF state. In addition to the photovoltaic effect, while in the ON state,
the reduced contact resistance allowed the photo-thermoelectric effect to contribute
to the photocurrent. The Seebeck coefficient of few-layer BP was estimated to be
≈100 V K–1 at 77 K, which reduces to ≈5 V K–1 with negative gate voltages. In
addition, operating with the photovoltaic effect in the OFF state, the device showed
a strong polarization-dependent photocurrent response, arising from the directional
dependence of the interband transition in the anisotropic band structure of BP.
Recently, Wu et al. examined the optoelectronic properties of few-layer-BP-based
photodetectors over a wide spectrum from UV to NIR (310 to 950 nm).[218]
Interestingly, the device demonstrated two distinct photoresponse behaviors in the
low-energy range (visible–NIR region, 400–950 nm) and high energy range (near-
UV region, below 400 nm). As shown in Figure 6b, the responsivity reaches its
highest value of 9×104 A W–1 in the near-UV region and decreases by several orders
of magnitude, down to ≈1.82 A W–1, in the visible–NIR region. The corresponding
response speed changes from tens of seconds to several milliseconds in the high-
energy and low-energy wavelength regions. The detectivity achieved by this device
in the near-UV region is as high as ≈ 3×1013 Jones. This colossal UV responsivity
is attributed to the resonant interband transition between two specially-nested
valence and conduction bands, which provides an unusually high density of states
for highly efficient UV absorption due to the singularity in the joint density of
states.
The responsivity of few-layer BP phototransistors can be further improved by
lowering the working temperature and scaling down the channel length. For
example, Huang et al. have exfoliated few-layer BP flakes with thicknesses of ≈8
nm onto SiO2/Si substrates and defined source/drain electrodes with various
channel lengths, followed by the deposition of Ni/Au metal contacts.[219] The
photocurrent of the BP transistor with a 1 μm channel increased significantly with
decreasing working temperature from 300 K to 20 K in a broad wavelength range
in the p-channel of the transistor, giving rise to the highest responsivity, ~7×106 A
W–1 at 20 K, and a much lower responsivity of ~6.7×105 A W–1 at 300 K (Figure
6c). The increased responsivity was attributed to the improved carrier generation-
recombination rate and better carrier transport due to the freezing of interface and
bulk traps, as well as the reduced phonon scattering. Further scaling of the channel
length down to 100 nm led to a much-improved responsivity of up to ≈4×107 A
W–1 at 20 K and ≈4.3×106 A W–1 at 300 K (Figure 6c) due to a larger transverse
electric field and smaller carrier transit times in shorter channel devices.

Figure 6. a) Responsivity of a few-layer BP photodetector versus excitation


wavelength at a constant power (red squares). Reproduced with permission.[216]
Copyright 2014, American Chemical Society. b) Responsivity of a few-layer BP
phototransistor within the energy range 1–4 eV measured at VBG = 80 V and VSD =
0.1 V. Inset shows AFM scan across the channel area of the device showing the
thickness of the black phosphorus flake to be ≈ 4.5 nm. Reproduced with
permission.[218] Copyright 2015, American Chemical Society. c) Responsivity of
few-layer BP phototransistors with different channel lengths under the same
driving voltage (Vg = -20 V, Vd =1 V). Reproduced with permission.[219] Copyright
2016, Wiley-VCH. d) SEM image of a BP-based THz detector. e) Gate bias
FEATURE ARTICLE

dependence of the responsivity (RV) and photovoltage (𝛥u) at room temperature


for the BP THz detector. The red line (curve 1) was measured by impinging the
THz beam on the detector surface; the black line (curve 2) was measured while
blanking the beam with an absorber. Inset shows the antenna-coupled structure.
(d,e) Reproduced with permission.[221] Copyright 2015, Wiley-VCH.
High-performance few-layer-BP-based photodetectors have been used for imaging
applications. Engel et al. reported that a photodetector based on 120 nm few-layer
BP could acquire images with high resolution both in the visible and IR spectral
regions.[220] The device was operated with the photo-thermoelectric effect. The test
object was raster scanned with respect to a tightly focused laser beam with a
wavelength of 532 nm or 1550 nm, and the reflected light signals were received
by the BP photodetector and converted into electrical currents for imaging
formation.
Few-layer BP also shows great promise in THz detection. A room-temperature THz
detector based on 10 nm thick few-layer BP was recently reported.[221] The device
structure with asymmetric antennas-coupled few-layer BP FET is shown in Figure
6d. A few-layer BP flake was mechanically exfoliated onto a SiO2/Si substrate as
the active layer of the FET. Source- drain contacts patterned by e-beam lithography
and metal deposition were oriented along the armchair (x) BP axis. The channel
length was ≈2.7μm and the gate length was ~580 nm. An 80 nm-thick SiO2
insulating layer was deposited on the sample by sputtering, followed by patterning
of the gate electrode on the oxide. The source and gate electrodes were in the shape
of two halves of a planar bow-tie antenna, with a total length of 500 μm and a flare
angle of 90 ° for coupling THz radiation (inset in Figure 6e). By keeping the source
electrode grounded, the photoresponse (voltage response) was recorded at the drain
electrode in an open circuit configuration through a lock-in amplifier connected
to a low-noise voltage preamplifier. The detection mechanism was attributed to
the combination of the rectification of THz-induced overdamped plasma waves in
the BP channel and the photo-thermoelectric effect. The generated voltage
response was sensitive to both the frequency of the impinging radiation and the
gate voltage. The detector reached a maximum voltage response of ≈0.9μV, giving
rise to the maximum responsivity of 0.15 V W-1 under 0.298 THz radiation at
room temperature (Figure 6e). Moreover, a lower limit for NEP, ≈40 nW Hz-1/2, has
been achieved, which is comparable to that of the best bilayer graphene-based THz
photodetectors.
However, despite these aforementioned achievements, a huge challenge that
impedes the application of few-layer BP in electronics/optoelectronics is its
environment-induced degradation.[26] Exfoliated flakes are highly hygroscopic and
tend to condense moisture from the air on their surface, which likely leads to the
degradation of BP. Therefore, it is necessary to develop passivation or encapsulation
techniques in the future for practical applications.
3.4. Hybrid Phototransistors
Ultrasensitive photodetectors operating at very low light intensities require a gain
mechanism by which one incident photon can induce multiple carriers to conduct
the current in the devices. A promising strategy is based on hybrid phototransistors
that function with the photogating effect, where light is efficiently absorbed by
appropriate sensitizing centers while photogenerated electrons or holes can transfer
into conducting channel and recirculate many times within the lifetimes of the
opposite charges.[77,78] According to their work principles, hybrid phototransistors
based on 2D layered semiconductors can be classified into two types. The first is
based on devices in which 2D layered semiconductors only serve as light absorbing
media. In these devices, graphene is normally employed as the conducting
channel material for its high carrier mobility (hence a low carrier transit time) and
strong sensitivity of its conductance to electrostatic perturbation by adjacent
photocarriers. Another type of hybrid phototransistor contains active materials, i.e.,
2D layered semiconductors serve as both light absorbers and conducting media
while other sensitizers are used to enhance light absorbance or to extend the
detection range. The performances of these devices are summarized in Table 3.
Table 3. Performance of hybrid phototransistors based on 2D layered
semiconductors. ML: Multilayer, Gr: Graphene, FL: Few-layer, IGZO: Indium
gallium zinc oxide, QDs: Quantum dots, CIS: Copper indium sulfide.

Examples of light-absorbing media in the first type of hybrid device are MoS2,[222–
227]
Bi2Te3,[228] and InSe.[229] Due to their strong light absorption and tunable
bandgaps, these devices are capable of achieving high gains with a detection range
spanning from the UV to NIR region. In 2013, Roy et al. reported a hybrid
phototransistor that consists of a monolayer graphene over a multilayer MoS2, both
of which were exfoliated from their bulk crystals, respectively.[222] A very high
responsivity of 5×108 A W–1 was obtained in the devices at room temperature and
increased to 1×1010 A W–1 at 130 K (Figure 7a), while the response times were very
slow (approximately tens of seconds). The device also demonstrated an ultrahigh
gain of ≈4×1010. The high performance is attributed to several different processes.
Under illumination, photoexcited holes generated inside the MoS2 are trapped by
localized states in the MoS2, while photoexcited electrons are transferred to the
graphene channel under the influence of a gate electric field (Inset in Figure 7a,
top-right panel). These electrons recombine with holes induced by the negative gate
bias in the channel and increased channel resistance, giving rise to a large net
photocurrent. Meanwhile, the trapped holes reside longer in the MoS2 and act as a
local gate, leading to strong photogating effect on the graphene channel through
capacitive coupling. The long lifetimes of holes were also responsible for the large
response times. Afterwards, Zhang and co-workers demonstrated a hybrid
photodetector composed of a CVD-grown single-layer MoS2 covered by a CVD-
grown graphene monolayer, which also showed a high responsivity of ≈1.2×107 A
W–1 and a gain up to 108.[223] Similarly, the device showed very slow response times
ranging from tens to hundreds of seconds. Compared with the value reported by
Roy et al.,[222] the relatively lower responsivity achieved in this work is likely due
to the lower carrier mobility in CVD graphene. However, CVD graphene and MoS2
are more conveniently prepared with large area and are therefore more suitable for
practical applications. As shown in Figure 7b, the Ids–Vg curves shift horizontally
with increasing incident powers, which undoubtedly confirms that the sensing
mechanism is dominated by the photogating effect. It is noteworthy that the device
performance decreases significantly when measured in high vacuum, which is
attributed to the decreased effective electric field at the graphene/MoS2 interface
due to the desorption of adsorbates and/or charged impurities in vacuum.
Graphene covered with other 2D layered materials such as Bi2T3[228] and InSe[229]
also exhibited a pronounced photogating effect. However, the devices showed
inferior responsivities between 35 A W–1 and 940 A W–1. In particular, Qiao et al.
reported a hybrid photodetector that consists of monolayer graphene covered with
epitaxially-grown B that consists of monolayer graphene covered with epitaxially-
grown Bi2Te3 nanoplates. The device shows an obvious photoresponse with a
detection range extending to the NIR (980 nm) and a telecommunication band
(1550 nm) (Figure 7c) due to the small bandgap of Bi2Te3.[228]
In the second type of hybrid device, the sensitizers that are employed to enhance
light absorbance or to extend the detection range are normally quantum dots
(QDs),[230–233] organic molecules,[234,235] and CH3NH3PbI3 perovskite.[236,237] For
example, Kufer et al. reported a hybrid photodetector composed of bilayer MoS2
covered with 1,2-ethanedithiol ligands and (EDT)-capped colloidal PbS QDs.[230]
The short ligands capped on the surface of the QDs are crucial to the photodetection
performance due to the improved charge transport of the QDs’ film and charge
transfer between the QDs and MoS2. Benefiting from the strong, size-tunable light
absorption in the QDs and high carrier mobility in the MoS2, the hybrid photo-
detector demonstrated a high responsivity of up to 106 A W–1, which surpassed that
of photodetectors based solely on MoS2 or PbS by several orders of magnitude.
Moreover, the detection range can be extended to the short-wave infrared (SWIR)
region, which is far beyond the detection limit of pure MoS2. Another advantage is
its decreased NEP and consequently increased detectivity of up to 7×1014 Jones,
achieved by suppressing its dark current by tuning the transistor in the OFF state
(Figure 7d). The photodetection mechanism also relies on a strong photogating
effect. Light absorption in the PbS QDs generates electron–hole pairs, which are
separated by the electric field at the PN interface between PbS and MoS2. Holes
are trapped within the QDs’ layer and increase the carrier lifetimes. Meanwhile,
electrons are transferred to the MoS2 and circulate through the channel before
recombination, leading to a high responsivity. More recently, Kang and co-workers
demonstrated a hybrid phototransistor comprising a CH3NH3PbI3 perovskite film
on top of APTES-doped MoS2.[237] Due to high absorption and long carrier
diffusion lengths of the perovskite, a large number of photogenerated electrons and
holes can be transferred to the MoS2. This effect, combined with reduced
photocarrier recombination and increased carrier mobility in MoS2 induced by
electron doping, gives rise to a significantly enhanced responsivity as high as
1.94×106 A W–1 and a detectivity of 1.29×1012 Jones.
In summary, hybrid phototransistors based on 2D layered semiconductors that
make use of the photogating effect show large responsivities and gains that are
normally several orders of magnitude higher than those of devices composed of
2D layered semiconductors alone (Table 3). Moreover, their operation can be
extended to the NIR or even the SWIR region by sensitizing appropriate light
absorbers. However, these devices usually show relatively long response times,
making them more suitable for special applications where speed is not a necessity.
Compared to hybrid devices with a graphene conducting channel, devices
employing other 2D layered semiconductors with intrinsic bandgaps as conducting
materials can generate a lower dark current and thus larger detectivities, which
might be more suitable for practical applications.

Figure 7. a) Responsivity of a graphene/MoS2 hybrid device as a function of


illumination power. Dashed line is a fit of the form A1/(A2+PLED), considering A1
and A2 as the fitting parameters. Top-right inset: schematic of charge exchange
process for Vg<<VT. Bottom-left inset: comparison of low-temperature and room-
temperature responsivity. Reproduced with permission.[222] Copyright 2013, Nature
Publishing Group. b) Transfer curves for a graphene/MoS2 photodetector under the
exposure of light with various powers. Inset shows the schematic illustration and
an optical image of the photodetector based on graphene/MoS2 stacked layers.
Reproduced with permission.[223] Copyright 2014, Nature Publishing Group. c)
Photocurrent of a graphene/Bi2Te3 heterostructure device as a function of
photoexcitation wavelength (from 400 to 1550 nm). The green dots are
experimental data, and the blue line is a visual guide. Inset shows schematic of the
heterostructure phototransistor device. Reproduced with permission.[228] Copyright
2015, American Chemical Society. d) Drain current IDS (under dark conditions),
responsivity R, and the shot-noise limited detectivity D* of a MoS2/PbS QDs
hybrid photodetector as a function of back-gate voltage VG measured at VDS = 1
V. Reproduced with permission.[230] Copyright 2015, Wiley-VCH.
4. Photodiodes Based on 2D Layered Semiconductors
The photodiodes based on 2D layered semiconductors have been realized with
different structures and principles. Table 4 summarizes the figures of merit of some
representative photodiodes based on 2D layered semiconductors and artificial
heterostructures.
4.1. Local Electrostatically Defined PN Junctions
The ultimate thickness of 2D layered semiconductors allows tuning their carrier
densities, band alignment, and polarities via electrostatic doping[25,48,49] or
chemical doping.[50–52] There exists a variety of 2D layered semiconductors, such
as WSe2,[48] MoSe2,[49] and few-layer BP,[25] that can readily exhibit ambipolar
transport properties and enable PN junctions by local electrostatic gating. In this
section, we introduce the studies related to photodiodes based on such PN
junctions.
A typical schematic illustration of a local electrostatically defined PN junction is
shown in Figure 8a. The fabrication of such PN junctions usually begins with standard
electron-beam lithography and metal deposition to define split gate electrodes on a
cleaned SiO2/Si substrate, followed by covering with a dielectric layer, which can
be Si3N4,[238] HfO2,[239] or layered h-BN.[240–243] Afterwards, single-layer WSe2 or
few-layer BP is mechanically exfoliated from their bulk crystals (multilayered
MoSe2 is synthesized by a CVT method[243]) and transferred onto the substrate, and
source and drain contacts are defined by another electron-beam lithography and
metal deposition process. In particular, metals with high and low work functions
can be chosen to facilitate hole and electron injection, respectively.
Table 4. Performance of photodiodes based on 2D layered semiconductors and
artificial heterostructures. FL: Few-layer, AuNPs: Au nanoparticles, ML:
Multilayer, Gr: Graphene, C8-BTBT: Dioctylbenzothienobenzothiophene,
SWNTs: Single-walled carbon nanotubes.
In 2014, three research groups reported local electrostatically-defined PN junctions
based independently on single-layer WSe2.[238–240] The local gates can effectively
control the charge carrier types and concentrations in the WSe2 channel. Thus
gate biases with opposite polarity can induce different carrier types in the adjacent
parts of the channel, which leads to a rectifying junction (PN or NP depending on
gate bias polarities), as identified by the rectifying Ids–Vds curves in the dark, shown
in Figure 8b. The curves all demonstrate high rectification factors with large shunt
resistances and low reverse saturation cur- rent, indicating high-quality PN
interfaces. It is important to note that the device can be operated as a resistor (PP
or NN) by applying gate biases with the same polarity (Figure 8b). These PN
junctions can be explored with diverse functionalities, such as photodiodes,
photovoltaic devices, and light-emitting diodes. Here, we only focus on their
application as photodiodes.
Under illumination, the Ids–Vds characteristics of a device introduced by Zaumseil
et al.[238] are shifted down (or up) when biased in PN (or NP) diode configuration,
giving rise to a short- circuit current ISC and an open circuit voltage VOC (Figure 8c).
This is clear evidence for the photovoltaic effect expected in the PN junction. Figure
8d depicts a photocurrent map obtained with SPCM of another device in NP diode
configuration.[239] It is obvious that the photocurrent is symmetric and centered at
the depletion region (separation between the gates), which strongly suggests that
the photoresponse is dominated by the PN junction rather than the Schottky barriers
at the contacts. The Ids–Vds curves under varied incident powers are shown in Figure
8e. The photocurrent at Vds = -2.0 V increases linearly with incident power up to
8.0 μW (inset in Figure 8e), resulting in a responsivity of 210 mA W-1. The
maximum EQE is estimated to be 0.2% under 522 nm illumination.
The PN photodiode can also work at zero-biased Vds condition with good
performance, as demonstrated by Groenendijk et al.[241] The ISC rises linearly while
the VOC increases logarithmically with increasing optical power when the device is
biased in PN diode configuration, as shown in Figure 8f. This also confirms that the
dominant photocurrent generation mechanism is the photovoltaic effect as well as
ideal diode behavior. Figure 8g shows the comparison of the time-dependent photo-
current of this device at constant Vds = 0 V or -1.0 V, from which a larger photocurrent
of ≈0.43 nA at Vds = -1.0 V is observed, corresponding to a maximum responsivity of
0.70 m A W–1. At zero-biased Vds, the response times are ≈12 ms, which decrease
in reverse bias. The EQE estimated from ISC reaches a maximum of ≈0.1%. It is
interesting to note that when the device is biased in PP configuration, the
photocurrent originates from the photo-thermoelectric effect, as identified by the
2D gate photocurrent maps obtained under above-bandgap and below bandgap
illuminations. This can be explained by the presence of a temperature gradient,
which induces a photo-thermoelectric voltage that can drive a current in the device
due to strong light absorption in the local thicker part of the flake.
In addition to WSe2, such PN junctions have also been realized with few-layer
BP[242] or multilayered MoSe2.[243] For example, Buscema et al. reported a few-
layer-BP-based PN junction, employing a thin h-BN flake as an atomically-flat and
disorder-free dielectric layer on two split gates.[242] Benefiting from the
ambipolarity, local electrostatic gating can effectively control the charge carrier
type and density in few-layer BP. As shown in the dark Ids–Vds curves (Figure 8h),
the device readily demonstrates rectifying behavior when the two gates are biased
with opposite polarity (PN or NP configuration), while the curves show almost
linear behavior when the two gates are biased with the same polarity (PP or NN
configuration). Under illumination, the photocurrent and photovoltage are only
generated when the two gates are oppositely biased, indicating that the
photoresponse generation mechanism relies on the photovoltaic effect originating
from the PN or NP junction in the channel. With increasing incident power, the Ids–
Vds curves shift down towards more negative current values in PN configuration,
giving rise to increased ISC and VOC. Interestingly, the photoresponse can be
extended to the NIR spectrum (Figure 8i) due to the narrow bandgap of few-layer
BP. The device reaches a maximum responsivity of 28 mA W–1 and decreases to
several mA W–1 as excitation wavelength increases. The maximum EQE is ≈0.1%
at 640 nm illumination and the response times are ≈2 ms.
Figure 8. a) Schematic illustration of a local electrostatically defined PN junction
based on single-layer WSe2. Ids–Vds characteristics of the PN junction device
measured in varied gate biasing configurations b) in darkness and c) under optical
illumination, respectively. Reproduced with permission.[238] Copyright 2014,
Nature Publishing Group. d) Photocurrent map of the PN junction device in NP
configuration. e) Ids–Vds characteristics of the device in NP configuration under
various laser powers. Inset shows Iph at Vds = -2 V as a function of laser power. (d,e)
Reproduced with permission.[239] Copyright 2014, Nature Publishing Group. f) Voc
and Jsc of the PN junction device in PN configuration as a function of incident light
power intensity. g) Time-dependent photocurrent of the device at Vds = 0 V and –1
V in PN configuration, respectively. Inset shows Ids–Vds characteristics in darkness
and under illumination. (f,g) Reproduced with permission.[241] Copyright 2014,
American Chemical Society. h) Ids–Vds characteristics of the PN junction device
measured in varied gate biasing configurations in dark. i) Ids–Vds characteristics of
the device in PN configuration measured in darkness and under NIR illuminations.
h,i) Reproduced with permission.[242] Copyright 2014, Nature Publishing Group.
In summary, thanks to their ambipolar property, some 2D layered semiconductors
can form lateral PN junctions with local electrostatic gating. In the PN junctions,
photocarrier generation, separation, and transport processes can be controlled by
the external electric fields applied on the split gates. With their unique optical and
electrical properties, PN junctions composed of single-layer WSe2 or few-layer BP
have been explored as photodiodes, showing responsivities between 0.70 mA W–1
to 270 mA W–1 and EQE values in the range of 0.1–0.2%. The response times
range between several to tens of milliseconds, which are comparable to the fastest
phototransistors made from 2D TMDs or few-layer BP. The NEPs of these
photodiodes are expected to be low due to the reduced dark current. Devices can
be further optimized in their structure to increase light absorption or decrease
contact resistance.
4.2. Graphene/2D Layered Semiconductor/Graphene Heterostructures
Rapid advancements in isolation and deterministic transfer of various 2D layered
materials[244,245] allow the possibility of integrating them in vertical stacks for novel
electronic and optoelectronic devices. In this section, we introduce the
photodetection applications of graphene/2D layered semiconductor/graphene
vertically stacked heterostructures. A schematic illustration of a typical
heterostructure is shown in Figure 9a. In the device, carrier transport occurs in the
2D layered semiconductor in the out-of-plane direction, while the graphene layers
serve as trans- parent charge extraction contacts. The finite density of states in
graphene and weak electrostatic screening effect allow effective modulation of the
Fermi level of graphene, and thus Schottky barrier height, between graphene and
the semiconductor by electrostatic gating or chemical doping. Photocarrier
generation, separation, and transport processes within the heterostructures can be
modulated by the gate as well.

Figure 9. a) Schematic illustration of a graphene/MoS2/graphene vertically stacked


heterostructure. Reproduced with permission.[246] Copyright 2013, Nature
Publishing Group. b) Ids–Vds characteristics of a graphene/WS2/graphene vertical
heterostructure as a function of gate bias. The right (left) y-axis gives the current
magnitude for measurements: in darkness, green traces; under illumination, blue to
red traces. c) Schematic band alignments in equilibrium (top panel) and under gate
bias or chemical doping (bottom panel). (b,c) Reproduced with permission.[39]
Copyright 2013, American Association for the Advancement of Science. d)
Schematic illustration of the experimental setup of the time-resolved photocurrent
FEATURE ARTICLE

measurement. e) Photoresponse rate and time constant of the device as functions of


layer thickness of the WSe2. f) Photoresponse rate and time constant of the device
as functions of working bias. d-f) Reproduced with permission.[247] Copyright 2015,
Nature Publishing Group.
In 2013, Britnell et al. reported a vertically stacked heterostructure composed of
WS2 with a thickness of 5-50 nm sandwiched between two graphene layers.[39] In
the dark, the device acts as a tunnelling transistor where the current is modulated
by the gate electric field, as shown in Figure 9b (the green traces). Under
illumination, the device shows a pronounced photoresponse with strong gate-
tunable photocurrent (Figure 9b, blue to red traces). The I–V characteristics show
linear behavior at low bias and become non-linear (current saturation) at high bias
due to the finite number of free carriers in the photoactive region. The photocurrent
generation is based on the photovoltaic effect, which can be understood with the
help of the band diagrams depicted in Figure 9c. Ideally, the band alignment
between the graphene layers and WS2 is symmetric; therefore, photogenerated
electrons/holes in WS2 have no preferred diffusion direction, giving rise to zero net
photocurrent (Figure 9c, top panel). However, a built-in electric field across the
WS2 can be formed due to a difference in doping levels or electrostatic gating in
graphene layers, which can rapidly separate photogenerated electrons and holes and
results in net photocurrent (Figure 9c, bottom panel). The gate-tunable built-in
electric field thus allows the heterostructure to function as a photodiode or a
photovoltaic cell. The heterostructure reaches the highest EQE above 30% and
decreases quickly with increasing incident power. This is likely due to the
absorption saturation in WS2 as well as the screening of the built-in electric field
by excited electrons in the conduction band of WS2. Further integration with
plasmonic metal nanostructures can lead to a 10-fold improvement in the
photocurrent (the responsivity increases from ≈10–2 A W–1 to ≈10–1 A W–1).
Yu et al. demonstrated a similar heterostructure comprising 50 nm thick MoS2 as
the photoactive layer.[246] The heterostructure shows a strong photoresponse
controllable by the gate voltage and fast response times on the order of ≈50 μs. Its
EQE reaches a maximum value of ≈25% and depends significantly on the incident
light intensity and wavelength. Moreover, the polarity and amplitude of
photocurrent can be modulated to an even higher extent by effectively changing the
direction and strength of the built-in electric field in a dual-gated graphene/
MoS2/graphene heterostructure. A higher EQE of up to 55% has been achieved in
an asymmetrical vertical structure of graphene/MoS2/metal (Ti), where MoS2-Ti
forms nearly ohmic contact while graphene-MoS2 contacts form a Schottky
junction, and the responsivity of the device reaches ≈0.22 A W–1. In both studies,
photocurrent maps via SPCM measurements reveal that the photocurrent is
primarily generated in the regions of the heterostructures with asymmetric
potentials.
Recently, Massicotte et al. have investigated the dynamics of the photoexcited
charges in a vertically stacked graphene/WSe2/graphene heterostructure via time-
resolved photocurrent measurements by exciting the device with a pair of
ultrashort pulses (≈200 fs) separated by a variable time delay (Figure 9d).[247] They
found that the fastest intrinsic response time can reach ≈5.5 ps in heterostructures
based on mono- layer or tri-layer WSe2. The response time increased to several
nanoseconds for ≈40 nm thick WSe2-based heterostructure (Figure 9e) and depends
strongly on the working bias (Figure 9f). Limited by the instruments and their
resistance- capacitance (RC) time, the heterostructure composed of WSe2 with a
thickness less than 10 nm shows a real-time photoresponse time less than ≈1.6 ns,
which is significantly faster than other TMDs based photodetectors. More
importantly, the EQE value of the device comprising tri-layer WSe2 reaches 7.3%,
corresponding to a high IQE value over 70%. In addition to TMDs, another 2D
layered chalcogenide, InSe, has also been integrated with graphene in such a
heterostructure.[248] Working at forward bias, the device shows a maximum
responsivity of ≈105 A W–1 and a detectivity of ≈1015 Jones at low laser power; both
decrease rapidly, however, with increasing laser power.
In summary, vertically stacked graphene/2D layered semiconductor/graphene
heterostructures demonstrate tunable built-in electric fields across their
heterostructures by electrostatic gating or chemical doping. This is a major
advantage over conventional photodiodes. Consequently, photocarrier generation,
separation, and transport processes can be readily manipulated. The photodetectors
based on such heterostructures can reach responsivities between ≈10–2 A W–1 to
≈105 A W–1, EQE values in the range of 7.3%–55%, and fast response times
between ≈1.6 ns to ≈50 μs. Their performances are superior to those obtained from
local electrostatically-defined PN junctions based on 2D layered semiconductors.
This is likely due to the atomically thin carrier transient paths in their vertical
heterostructures that facilitate efficient charge separation and collection.
4.3. 2D Layered Semiconductor/2D Layered Semiconductor Heterostructures
In the previous section, we discussed graphene/2D layered
semiconductor/graphene vertically stacked heterostructures and their
photodetection applications. In this section, we introduce the photodetection
applications of 2D layered semiconductor/2D layered semiconductor
heterostructures, including vertically stacked and laterally stitched ones. In general,
vertically stacked heterostructures can be realized either by mechanical stacking,
where the stacking order and the interface of the heterojunctions are critical to their
electrical/optoelectronic properties, or by direct synthesis via CVD methods. On
the other hand, laterally stitched heterostructures can only be grown by direct
synthesis, since 2D layered materials are covalently bonded laterally. For more
details about the preparation of 2D layered materials-based heterostructures,
interested readers should refer to related perspectives and review papers.[44,45]
4.3.1 Vertically Stacked Heterostructures
A vertically stacked heterostructure based on single-layer 2D semiconductors
with different conductive types can form a heterojunction usually with type II
band alignment, which facilitates efficient electron–hole separation for light
detection or harvesting.[249,250] Compared to conventional heterostructures, these
heterostructures should in principle possess many distinct advantages such as
atomically sharp interfaces, no interdiffusion of atoms, digitally controlled layered
components, and no lattice parameter constraints. Strong interlayer coupling of
charge carriers[249] and ultrafast photoexcited charge transfer within 50 fs to sub-
picosecond at the interface of single layers[250–252] have been observed in such
heterostructures, enabling the design of novel devices for optoelectronics and light
harvesting.
[253–256]
Recently, TMD/TMD vertical heterostructures, including WSe2/MoS2,
[257,258] [259]
WS2/MoS2, WSe2/WS2, WSe2/MoSe2,[260] and α-MoTe2/MoS2[261]
have been widely studied in the field of photodetection. For example, Lee et al.
reported a vertical heterostructure composed of WSe2 and MoS2 single layers
crossed on top of each other, as shown in Figure 10a.[253] Palladium and
aluminum contacts were chosen for p-WSe2 and n-MoS2 to facilitate holes and
electrons injection, respectively. In darkness, the heterostructure shows rectifying
behavior, which can be tuned by gate voltages. Under illumination, the device
shows a strong gatetunable photovoltaic response, as evidenced by the Ids–Vds
characteristics in Figure 10b. A maximum photoresponse is observed at Vg = 0 V,
the position where the heterostructure exhibits the strongest rectifying behavior. A
spatial photocurrent map at Vds = 0 V indicates that the photocurrent is generated
primarily in the WSe2/MoS2 junction area (inset in Figure 10b). This result,
combined with PL measurements, reveals that the photovoltaic response
originates from the rapid separation of photocarriers between the two TMD layers.
In particular, the strong gate-dependent photoresponse can be related to the
interlayer tunnelling-mediated recombination of the majority carriers across the
interface of the heterostructure, which can be tuned by gate voltages. A maximum
responsivity of ≈2 mA W–1 can be further increased to ≈10 mA W–1 by using
graphene electrodes, which allows carrier collection by direct vertical charge
transfer and thus reduces interlayer recombination losses of photocarriers. The
responsivity can be improved even further to ≈120 mA W–1 in the heterostructure
consisting of two multilayer TMDs due to increased light absorption. The EQE also
increases with increasing layer thickness and reaches the highest value of ≈34%
in the multilayer heterostructure.
Very recently, Pezeshki and co-workers fabricated a vertical heterostructure
consisting of few-layer α-MoTe2 and MoS2 and probed its electronic and
optoelectronic properties.[261] Their results show that the heterostructure exhibits a
strong photovoltaic response from the visible to NIR (800 nm) range. The
responsivity and EQE reach the maximum values of 322 mA W-1 and 85% under
visible light and decrease rapidly to 38 mA W–1 and 6% at 800 nm NIR light,
respectively. The response times are on the order of ≈25 ms, and the obvious
NIR response is likely attributed to the small bandgap of few-layer α-MoTe2.[262]
In particular, broadband photodetection with a response range up to 2400 nm can be
realized by sandwiching graphene with a gapless band structure that enables
broadband light absorption in a WSe2/MoS2 vertical heterostructure (inset in Figure
10c).[263] Similarly, this heterostructure shows an apparently gate-tunable rectifying
behavior and a photovoltaic response. The responsivity and detectivity are as high as
104 A W-1 and 1015 Jones at visible wavelength and decrease drastically to 10-1 A W–
1
and 109 Jones at IR wavelength (2400 nm) at forward bias, respectively (Figure
10c). The response times are tens of microseconds, and the broadband wavelength-
dependent photoresponse can be attributed to the energy band diagrams of the
heterostructure. In the visible range where photon energy is larger than the
bandgaps of the TMDs, all three layered materials (WSe2, graphene, and MoS2) can
absorb light efficiently and produce abundant electron–hole pairs, resulting in a
considerably larger photocurrent (Figure 10d, left panel). However, under IR light
where photon energy is smaller than the bandgap of WSe2, light can only be
absorbed by graphene and generates electron–hole pairs (Figure 10d, right panel),
which leads to a much smaller photocurrent in the IR range.
Moreover, vertical heterostructures consisting of few- layer BP and single- or
few-layer MoS2 have recently been explored.[264,265] As expected, these
heterostructures show obvious current-rectifying behaviors and photovoltaic
responses that are tunable by gate. The device reaches a maximum responsivity of
3.54 A W–1 at forward bias, which decreases to 418 mA W–1 at reverse bias.[264]
Very recently, a broadband photodetector with a visible to NIR detection range has
also been achieved in such a heterostructure.[266] The strong response in the NIR
is attributed to the photocarrier excitation in the few-layer BP due to its narrow
bandgap. The responsivity of the heterostructure is 22.3 A W-1 under 532 nm
illumination and decreases to 153.4 mA W–1 under 1550 nm illumination.
In addition to the vertical heterostructures described, those 2 comprising TMDs
[267] [268]
and/or other 2D materials, such as InAs/WSe2, SnS/SnS2, GaTe/MoS2,[269]
[270] [271]
C8-BTBT/MoS2, and CH3NH3PbI3/WSe have also demonstrated diodelike
rectifying behaviors and pronounced gate-tunable photovoltaic responses. In
particular, Cheng et al. reported vertically stacked heterostructures composed of
organohalide perovskites (CH3NH3PbI3) and 2D layered materials.[271] Two device
architectures have been realized, including a graphene/CH3NH3PbI3/graphene
vertical stack and a graphene/WSe2/ CH3NH3PbI3/graphene vertical heterojunction.
2D CH3NH3PbI3 is obtained by converting ultrathin PbI2 layers exfoliated from a
PbI2 crystal via a vapor phase intercalation method. The first device reaches a
maximum responsivity of ≈950 A W–1 and a photoconductive gain of ≈2200- much
higher than those of perovskite photoconductors with lateral configuration.[272,273]
Response times are on the order of several millimeters, and the high performance
can be attributed to the extremely short vertical carrier transit path and thus a small
carrier transit time. The second device shows diode-like rectifying behaviors within
a positive VG range, whereas nearly symmetric I–V characteristics can be observed
within the negative VG range (Figure 10e). This is attributed to the transition from
the PN to PP junction at the perovskite/WSe2 interface due to the ambipolar nature
of WSe2 under different gate biases. Under illumination, the heterostructure
demonstrates a strong gate-modulated photovoltaic response, as evidenced by the
I–V characteristics in Figure 10f.
Figure 10. a) Bottom left: Schematic diagram of a van der Waals-stacked MoS2/WSe2
heterojunction device. Top: enlarged crystal structures of MoS2 and WSe2 single layers.
Bottom right: Optical image of the fabricated device. b) Photoresponse characteristics of the
heterostructure at various gate voltages under white-light illumination. Inset shows the
photocurrent map of the heterostructure at Vds = 0 V. (a,b) Reproduced with permission.[253]
Copyright 2014, Nature Publishing Group. c) Responsivity R (left) and specific detectivity
D* (right) of a WSe2/graphene/MoS2 for wavelengths ranging from 400 to 2400 nm. Inset
shows a schematic illustration of the heterostructure. d) Schematic band diagrams of the
heterostructure in UV/visible range (left panel) and IR range (right panel). (c,d) Reproduced
with permission.[263] Copyright 2016, American Chemical Society. Ids–Vds characteristics of
the graphene/WSe2/CH3NH3PbI3/graphene vertical heterostructure e) in darkness and f)
under illumination at varied VG. Reproduced with permission.[271] Copyright 2016,
American Chemical Society.
4.3.2. Laterally Stitched Heterostructures
Recently, TMD/TMD 2 lateral heterostructures such as MoS2/ MoSe2,[274]
WS2/WSe2,[274] MoSe2/WSe2,[275] WS2/MoS2,[276] WSe2/MoS2,[277] and
[278]
WSe2/MoS2 have been successfully realized by one-pot or two-step CVD
methods. These heterostructures exhibit diode-like behaviors such as gate-
modulated current rectifying properties (Figure 11a) and photovoltaic responses
(Figure 11b). For example, a WS2/WSe2 lateral heterostructure provides a
pronounced photoresponse with response times less than 100 μs (inset in Figure
11b).[274] Photocurrent mapping reveals that the photocurrent is generated at the
lightly-doped WS2 and WS2/WSe2 interface regions (inset in Figure 11a),
indicating that the photoresponse originates from photocarrier generation and
separation at the heterojunction. In addition, Feng et al. reported a lateral
heterostructure composed of InSe/CuInSe2, where CuInSe2 was prepared by
partially transforming InSe through a simple solid-state reaction method.[279]
Apparently, the heterostructure shows robust rectifying behavior with a small
ideality factor. Due to its narrow bandgaps, the lateral diode can provide a
broadband photoresponse (254-850 nm) with the maximum responsivity of 8.4
A W–1 under UV light. The responsivity depends strongly on the incident light
wavelength and decreases with increased incident power.
In summary, 2D/2D layered semiconductor vertical and lateral heterostructures can
be easily prepared by mechanical stacking or direct synthesis methods. The rich
family of 2D layered semiconductors with different electrical and optical properties
makes it possible to realize multifunctional devices. Photodetectors based on these
heterostructures show broadband responsivities in a wide range from several mA W–
1
to 104 A W–1 and response times between tens of microseconds to tens of
milliseconds. Further work, especially on lateral heterostructures, is needed to both
explore fundamental study and develop practical applications.

Figure 11. a) Gate-tunable Ids–Vds characteristics of a lateral WSe2-WS2 heterojunction


PN diode. Inset shows the photocurrent map of the lateral heterojunction. b) Ids–Vds
characteristics of the lateral heterojunction in darkness and under illumination. Inset
shows the time- dependent photoresponse of the device. (a,b) Reproduced with
permission.[274] Copyright 2014, Nature Publishing Group.
4.4. 2D Layered Semiconductor/2D Layered Semiconductor Homojunctions
4.4.1. Vertical Homojunctions
A homojunction with continuous band alignment possesses small carrier trap sites
at the interface and can provide low contact resistance at the interface and high
carrier transport efficiency through the interface, which is beneficial for
optoelectronic applications. Vertically stacked PN homojunctions composed of 2D
layered TMDs such as MoS2 and MoSe2 were introduced recently.[280,281] For
example, Li et al. reported the realization of a vertical PN junction comprising
ultrathin MoS2 layers, where p-MoS2 and n-MoS2 were achieved via chemical
doping with AuCl3 and benzyl viologen as dopants, respectively.[280] Figure 12a
shows schematic illustration of a typical vertical homojunction. When functioning
as an FET, the device shows ambipolar carrier transport with an obvious current
hysteresis (Figure 12b), contributed by electrons and holes with lateral in-plane
transport along the MoS2 layers, as well as vertical interlayer tunneling. The device
can also be used as a phototransistor with a pronounced photoresponse under
illumination (Figure 12b). Interestingly, the MoS2 homojunction provides a clear
gate-tunable current rectifying behavior via modulation of the potential barrier or a
reversed current rectification via direct tunneling. Under illumination, such a
homojunction exhibits a strong photoresponse and can function as a
phototransistor at forward biases and a photodiode at zero bias. The responsivity
of the photodiode is 30 mA W-1. In another work, Jin and co-workers realized a
FEATURE ARTICLE

vertical homojunction consisting of few-layer p-type MoSe2 by Nb-doping and


few-layer intrinsic MoSe2 with inherent n-type behavior.[281] The homojunction
shows a typical gate-modulated rectifying characteristic with a large current
ON/OFF ratio and a small ideality factor, implying ideal diode behavior. The
homojunction exhibits a pronounced photovoltaic response, as shown in Figure
12c. The responsivity reaches a maximum of several amps per watt at forward
bias and tens to hundreds of milliamps per watt at zero and reverse biases (Figure
12d).
4.4.2 Lateral Homojunction
A lateral homojunction has been realized in few-layer MoS2 containing p-MoS2
chemically doped with AuCl3 and intrinsic n-type MoS2. Partially stacked 2D h-
BN was used as a protecting mask to isolate the underlying intrinsic MoS2 from
the dopant (Figure 12e).[282] Asymmetric contacts of Pd and Cr/Au were
incorporated, respectively, in p-MoS2 and n-MoS2 to facilitate the injection of holes
and electrons. The device exhibits a pronounced gate-modulated current rectifying
behavior in the dark and a photovoltaic response under illumination as well. As
depicted in Figure 12f, the transfer characteristics under illumination show two
distinct photoresponse regions, which are attributed to the gate-induced
accumulation of electrons (at positive VG) or holes (at negative VG) that reduces the
barrier height between the conduction or valence band edge of MoS2 and the
contacts. At zero VG, the device gave a responsivity and detectivity of 5.07 A W–1
and 3×1010 Jones at forward bias, respectively, and 0.33 A W–1 and 1.6×1010 Jones
at reverse bias, respectively. The response times are in the order of 100–200 ms. More
recently, Yu et al. reported a few-layer-BP-based lateral homo- junction with a similar
device geometry, where benzyl viologen (BV) was introduced as an effective electron
dopant in BP.[283] Due to the narrow bandgap of the few-layer BP, the device is capable
of detecting NIR light (1.47 μ m) with a responsivity of ≈180 mA W–1 and fast
response times on the order of tens of milliseconds.
Figure 12. a) Schematic illustration of a MoS2 vertical homojunction. b) The transfer

FEATURE ARTICLE
characteristics of the homojunction measured in darkness and under illumination during
both the forward and reverse sweeps. Reproduced with permission.[280] Copyright 2015,
Nature Publishing Group. c) Current density–voltage (JD–VD) curves of the MoSe2
homojunction PN diode in darkness and under light irradiation with various laser power
densities for VG = 0 V. Inset shows schematic illustration of the homojunction. d)
Responsivity and EQE for various VD as a function of light irradiation power. (c,d)
Reproduced with permission.[281] Copyright 2015, Wiley-VCH. e) Schematic illustration of
a MoS2 lateral homojunction. f) ID–VG curves of the lateral homojunction under light
exposure with different wavelengths. e,f) Reproduced with permission.[282] Copyright 2014,
American Chemical Society.
4.5. 2D Layered Material/Semiconductor Heterojunctions
Benefiting from deterministic transfer techniques[244,245] and advanced growth
methods,[4,22,24,26] 2D layered materials can be readily integrated with conventional
semiconductors to form van der Waals heterojunctions. This includes 3D bulk
materials such as Si,[284–293] GaN,[294] GaAs,[295,296] or 2D thin films such as ZnO
film,[297] CuPc film,[298] and rubrene single-crystal.[299] These heterojunctions with
atomically sharp interfaces can be exploited as optoelectronic devices with versatile
functionalities, including photodiodes, photovoltaic devices, and light-emitting
diodes.
Heterojunctions composed of 2D MoS2 and Si have been extensively studied as
photodetectors.[286–290] Schematic illustration of a typical device geometry is
shown in the inset of Figure 13a. A single-layer MoS2 flake is transferred on top
of a p-doped Si substrate, and a PN heterojunction forms at the Si/ MoS2 interface.
In darkness, the current–voltage (I–V) characteristic shows typical diode-like
behavior with a good rectifying factor (Figure 13a). Li et al. have studied the
photoresponse of both monolayer MoS2/n-Si and monolayer MoS2/p-Si
heterojunctions.[286] Both heterojunctions show obvious photoresponses at reverse
bias under illumination, which depend upon the incident light wavelength and
power intensity. The highest responsivity reaches 7.2 A W–1 for the MoS2/n-Si
heterojunction, while the value decreases to ≈1 A W–1 for the MoS2/p-Si
heterojunction. The underlying mechanisms of the different photoresponse
behaviors were further investigated by combining Kelvin probe microscope (KFM)
studies and band alignment analyses. The results show that the photodiode
behavior originates from the intrinsic built-in electric field at the MoS2/p-Si
interface, whereas the built-in electric field from a modulated barrier height and
width induced by photogenerated excitons in the vicinity of the MoS2/n-Si interface
is responsible. Heterojunctions can also be constructed by directly depositing
MoS2 films on Si substrates via magnetron sputtering techniques[288] or a dip-
coating method followed by an annealing process.[289] For example, Wang et al.
demonstrated that molecular layers of MoS2 can be deposited perpendicularly on an
Si substrate to form a PN heterojunction via a scalable magnetron sputtering
method, where the vertically standing layered structures offer high-speed paths for
the separation and transportation of photocarriers.[288] The strong light absorption
of the relatively thick MoS2 film in combination with the unique structure gives
rise to an outstanding photoresponse in a broadband wavelength range and robust
air stability. Working in photovoltaic mode (under zero-biased voltage), the device
shows a responsivity of ≈300 mA W–1, a detectivity as high as ≈1013 Jones due
to a depressed dark current, and response times of ≈3 μs.
The band alignment between MoS2 and Si unveils the working mechanism of the
heterojunction photodetectors. Li et al. studied the 3D band diagrams of an n-type
monolayer MoS2/p-type bulk Si heterojunction, a prototype of the 2D–3D
heterostructures.[290] The wavelength-dependent photocurrent measurements
strongly suggest type I band alignment between the heterojunction (Figure 13b),
where photocarriers can move along both the vertical (z) direction (e.g., flowing
from bulk Si into the MoS2) and in the lateral (x–y) direction (e.g., driven by an
in-plane electrical field). Moreover, the performance of the heterojunction is
mainly determined by the electron–hole pair excitation and collection in the MoS2
monolayer in the heterojunction.
In addition to MoS2, other 2D layered materials such as GaSe,[291] GaTexSe1–
[292]
x, and Bi2Se [293]3 have been integrated with Si to function as photodiodes. Xiu’s
group reported the realization of 2D GaSe/Si[291] and 2D GaTexSe1-x/Si[292] PN
heterojunctions via a layer-by-layer van der Waals epitaxial growth technique.
These heterojunctions demonstrate steady rectifying characteristics in the dark and
prominent photoresponse behaviors, which strongly depend on the incident laser
power, as expected in photodiodes (Figure 13c). In particular, the 2D GaSe/Si
heterojunction shows a maximum EQE of ≈23.6%, which rapidly decreases to
≈2% with decreasing GaSe layer number, presumably due to weakened light
absorption in thinner layers and the band structure transition.[291] The response
times are tens of microseconds. The fast response can be accounted for by the
strong built-in electric field induced by the highly depleted region in the GaSe,
which can efficiently separate photocarriers, as well as the short carrier transport
distance arising from the vertical heterostructure. The EQEs for the 2D GaTexSe1-
x /Si heterojunction reach a maximum value around 60% and decrease with
increasing Se composition, as shown in Figure 13d.[292]
More recently, Zhang et al. reported a 2D topological insulator Bi2Se3/Si
heterojunction by van der Waals epitaxial growth of Bi2Se3 films tens of nanometers
thick on a n-Si substrate via a PVD method.[293] The heterojunction routinely
demonstrates obvious diode-like characteristics and a strong light-intensity-
dependent photoresponse. The responsivity reaches 24.28 A W–1 at reverse bias,
which decreases to 2.60 A W–1 at zero bias. Moreover, this device also shows fast
response times on the order of several microseconds due to a built-in electric
field at the junction. Interestingly, the heterojunction shows an extended
photoresponse up to telecommunication wavelengths (1310 nm and 1550 nm),
which is related to the small bulk bandgap and gapless character of topological
surface states in Bi2Se3.
Figure 13. a) Current–voltage (I–V) characteristic of a p-Si/n-MoS2 heterojunction. Inset
shows schematic illustration of the heterojunction. Reproduced with permission.[285]
Copyright 2014, American Chemical Society. b) A cross-sectional schematic (x-z plane) of
the p-Si/n-MoS2 heterojunction (left panel), together with a band diagram of the 2D-3D
heterojunction along the x- and z-axes at zero bias under laser excitation. CB: conduction
band, and VB: valence band. Reproduced with permission.[290] Copyright 2015, American
Chemical Society. c) I–V characteristics of a 2D GaSe/Si heterojunction measured in
darkness and under illumination with various powers. Reproduced with permission.[291]
Copyright 2015, American Chemical Society. d) Composition-dependent photocurrent (left
panel) and EQE (right panel) of a 2D GaTexSe1-x/Si heterojunction under different laser
intensities. Reproduced with permission.[292] Copyright 2015, American Chemical Society.
e) Band diagrams of a 2D MoS2/ZnO heterojunction at reverse bias without and with an
applied pressure, which illustrate the photogenerated carriers and the piezophototronic effect
enhanced photocurrent. Reproduced with permission.[297] Copyright 2015, Wiley-VCH.
Apart from Si, monolayer MoS2 has been combined with GaAs to form
heterojunctions as photovoltaic-type detectors. In their recent study, Xu et al.
demonstrated that a MoS2/GaAs heterojunction shows a strong photoresponse in the
range of UV to visible light with a responsivity of 321 mA W–1, a detectivity of 3.5
×1013 Jones, and response times of tens of microseconds.[296] It is notable that the
responsivity can be increased by ≈ 30% to 419 mA W–1 via interface design by
inserting a h-BN layer as well as photoinduced doping of MoS2 by Si quantum
dots. Moreover, the heterojunction displays an increased detectivity as high as
1.9×1014 Jones due to the depressed dark current of the MoS2/h-BN/GaAs
sandwich structure, where the presence of h-BN raises the barrier height, leading
to increased shunt resistance and hence reduced thermal noise. More recently, Xue
and co-workers reported the realization of a 2D MoS2/ZnO film PN heterojunction,
where p-type MoS2 is achievable via a plasma-assisted doping method.[297] The
heterojunction shows an obviously gate-tunable rectifying behavior. Under UV
illumination, the device reaches the highest EQE of 52.7%, which can be enhanced
by more than four times via the piezophototronic effect. The detailed mechanism
for the enhanced photoresponse can be explained by the enlarged depletion region
within the ZnO film, which contributes to the generation and separation of
photocarriers. With external pressure applied to the ZnO film, the positive
piezoelectric charges yielded at the bottom surface of the ZnO lower the barrier at
the heterojunction interface and reduce the up-bending of the energy bands of the
ZnO, thus enlarging the depletion region within the ZnO film (Figure 13e).
Inorganic-organic heterojunctions combine the novel properties of inorganic solids
with those of organic materials, resulting in devices with diverse functionalities.
Recently, Liu et al. reported an inorganic–organic heterojunction composed of 2D
MoS2 and a rubrene single crystal that shows a gate-tunable rectifying
characteristic.[299] The heterojunction demonstrates a strong photoresponse with a
responsivity of 510 mA W–1 at forward bias and response times of several
milliseconds.
In summary, integration of 2D layered materials with conventional semiconductors
can lead to novel van der Waals heterojunctions with atomically abrupt interfaces,
harnessing the advantages of both materials. In general, photodetectors based on
these heterojunctions show responsivities less than 1 A W–1 at zero bias, which
can be increased to tens of amps per watt at reverse bias. The response times for
heterojunctions based on 2D layered materials and bulk Si are typically several to
tens of microseconds-much faster than PN junction photodetectors based on 2D
layered materials via local electrostatic gating. Moreover, the fabrication processes
are compatible with the mature Si technology, which might make such
heterojunctions more amenable to practical applications.
4.6 Other Artificial Heterostructures
Integration of a 2D layered semiconductor with two metal (semi-metal) contacts
possessing different work functions gives rise to differences in local doping of the
layer, and therefore an asymmetry in the Schottky barriers at the two contacts. Such
asymmetric Schottky barriers result in the formation of a built-in electric field,
which is highly desirable in a photodiode. Fontana et al. reported a multi-layer
MoS2 flake-based Schottky barrier photodiode, where Au and Pd asymmetric
contacts were chosen to achieve electron- and hole-doping, respectively (Figure
14a).[300] The diode demonstrates an obvious current rectifying behavior which is
expected in darkness (Figure 14b). Under illumination, a strong photovoltaic
response is observed, which is believed to originate from the photocarrier
generation and separation by the built- in electric field, as shown in Figure 14b.
The electronic and optoelectronic characteristics of such a heterostructure can be
improved by inserting an insulating layer between the semiconductor and metal
contact to form a metal-insulator- semiconductor (MIS) diode. In a recent study,
Jeong and co-workers presented an MIS diode consisting of graphene, h-BN, and
single-layer MoS2. It demonstrated improved cur- rent rectification and much
higher current flow over a metal-semiconductor (MS) diode and a PN junction
made of 2D TMDs, due to carrier tunnelling at forward bias and depressed carrier
tunnelling at reverse bias.[301] Moreover, the MIS diode exhibits a pronounced
photoresponse with a responsivity of 0.3 mA W–1 and response times of ≈10 s.
Heterostructures composed of 2D layered semiconductors with other
nanomaterials have also shown great promise in the field of photodetection. By
integrating p-type single- walled carbon nanotubes (SWCNTs) with n-type single-
layer MoS2, Jariwala et al. reported a vertically stacked PN junction diode with
electrical characteristics that can be modulated by the gate bias to achieve a wide
tunability of charge transport ranging from nearly insulating to highly
rectifying.[302] Under illumination, a strong photovoltaic response in a wide range
of irradiation wavelengths (500-1100 nm) is observed for the diode, indicating that
exciton and/or free carrier generation in both materials contribute to the
photocurrent. Photocurrent mapping via SPCM reveals that photocurrent
generation lies in the overlapped area of the two components, confirming that the
photocurrent is generated by the vertical heterojunction. Interestingly, as shown in
Figure 14c, the relative photocurrent contribution from the SWCNTs decreases
with decreased VG, which is explained by the shortened depletion region in
SWCNTs due to higher doping/majority carrier concentration induced by negative
VG on this side. The responsivity depends strongly on the gate bias and reaches
a maximum value of 0.1 A W–1 under 650 nm illumination, decreasing rapidly
under illuminations of longer or shorter wavelength. The diode shows fast response
times with an upper limit of 15 μs, also due to the extremely short distance the
photocarriers must traverse. Very recently, Lee and colleagues have demonstrated
a nanocrystalline graphene-MoS2 lateral interface heterostructure (Figure 14d),
which shows a distinct diode-like current rectifying behavior.[303] The rectifying
characteristic stems from built-in electric fields formed at the lateral interfaces and
a potential shift in graphene induced by the metallic edge effect of the MoS2. Under
illumination, the heterostructure shows a photoresponse with a photocurrent ≈500
times stronger than a vertically stacked graphene/MoS2 heterostructure. Moreover,
the lateral interface device shows response times of ≈2.8 s, also making it faster
than the vertical one (≈4.5 s). The improved photoresponse is attributed to the built-
in electric field, which can efficiently separate the photocarriers.
Figure 14. a) Optical image of an MoS2 Schottky barrier diode with asymmetric electrodes.
b) I–V characteristics of the Schottky diode in darkness and under illumination. Reproduced
with permission.[300] Copyright 2013, Nature Publishing Group. c) Photocurrent spectral
response of an SWCNTs/MoS2 PN heterojunction at varied VG. Reproduced with
permission.[302] Copyright 2013, National Academy of Sciences, USA. d) Photo image of a
nanocrystalline graphene/MoS2 heterostructure on a transparent flexible substrate. Inset
shows schematic illustration of the heterostructure. Reproduced with permission.[303]
Copyright 2016, Wiley-VCH.

5. Light-Matter Interaction for Sensitive Enhancement


Although 2D layered semiconductors demonstrate strong light-matter interaction, the
absolute light absorption is very low for such atomically thin layers. Further absorption
enhancement is highly desirable in 2D-layered-semiconductors-based optoelectronic
devices for some applications. Recently, several methods, including plasmonic
techniques, integration of optical waveguides, and microcavities have been developed
with the aim to increase light-matter interaction for light absorption enhancement. In
this section, we discuss techniques that are relevant to 2D-layered-semiconductor-based
photodetectors.
5.1. Plasmonic Nanostructures
Plasmonic metallic nanostructures can offer at least two ways to enhance light
absorption in 2D layered semiconductors. First, localized surface plasmons (LSPs) in
metallic nanostructures can lead to oscillation of conduction electrons in the
nanostructures, which greatly enhances local electromagnetic fields that improve light
absorption in semiconductors.[304,305] Moreover, the metallic nanostructures act as
subwavelength scattering elements that can couple and trap freely propagating plane
waves from the incident light into semiconductors.[305,306] In addition, plasmonic
nanostructures can provide another way to boost the photoresponses of detectors based
on 2D layered semiconductors via hot electrons injection that are generated by the
decay of the resonantly excited surface plasmons of the nanostructures.[307–309]
Normally, integration with metallic nanostructures can lead to a several-fold
improvement in responsivities for 2D-layered semiconductors-based photodetectors at
the wavelength range where plasmonic resonance occurs.[39,172,304–306] For example,
Miao et al. reported a two-fold enhancement of photocurrent upon depositing Au
nanoparticles sparsely onto a few-layer MoS2-based phototransistor under 632 nm
illumination.[305] A three-fold increase in photocurrent can be further attained by
utilizing periodic Au nanoarrays with stronger plasmonic resonance enhancement
(Figure 15a and b). The detection range of the photodetectors can also be extended to
wavelengths beyond the absorption edge of the 2D layered semiconductors with the
help of plasmonic enhancement. Recently, Wang and co-workers presented a bilayer
MoS2 photodetector with the photoresponse extended to NIR range by exploiting
plasmonic resonators.[309] The below-bandgap photocurrent originates from the
injection of hot electrons that are generated via the surface plasmon nonradiative decay
in the Au nanostructures. Consequently, the device demonstrates the highest hot-
electron-based responsivity of 5.2 A W–1 at 1070 nm, owing to trap-induced
photoamplification.

Figure 15. a) Schematic of a few-layer MoS2 phototransistor decorated with Au


nanoparticles. b) The photocurrent of the device as a function of illumination
wavelength. Reproduced with permission.[305] Copyright 2015, Wiley-VCH. c) Three-
dimensional illustration of a waveguide-integrated few-layer BP photodetector. d) The
normalized photoresponse as a function of light modulation frequency when BP is gated
to low and high doping, respectively. (c,d) Reproduced with permission.[313] Copyright
2015, Nature Publishing Group.
5.2. Optical Waveguides
Light-matter interaction enhancement can also be realized by integrating a 2D ultrathin
flake with an optical waveguide, where an evanescent field at the boundaries of the
waveguide can be absorbed by the flake to give rise to electron–hole pairs that
contribute to the photocurrent. In addition to graphene based photodetectors,[310–312] this
technique has been used to enhance the performance of a few-layer BP based
detector.[313] In this study, a multilayer BP flake is integrated onto a silicon photonic
waveguide with a few-layer graphene top gate (Figure 15c), which enables optimal
interaction with light and tunability of the carrier type and concentration. The
photocurrent generation mechanism is based on the photovoltaic effect for low-doped
BP, while it changes to the bolometric effect for highly n-doped BP. When the flake is
gated to be nearly intrinsic, the device with ≈11.5 nm BP reaches its optimal
photoresponse with an intrinsic (extrinsic) responsivity of 135 mA W–1 (18.8 mA W–1)
at telecommunication band (≈1550 nm). Importantly, due to the finite bandgap of few-
layer BP, the device exhibits a low dark current and gives rise to a depressed noise
level, a significant advantage over its graphene counterparts. In addition, the detector
shows a high response speed with a response roll-off frequency at 2.8 GHz for low
doping, which decreases to 0.2 MHz for high doping due to the change of photocurrent
generation mechanism, as shown by the photoresponse as a function of light modulation
frequency in Figure 15d.
5.3. Optical Microcavities
The integration of some 2D layered semiconductors with optical waveguides will not
amount to obvious performance enhancement because of their relatively large bandgaps
and thus negligible absorption in the telecommunication band. However, light-matter
interaction enhancement is achievable by integrating these semiconductors with optical
microcavities in the visible region. It has been recently reported that coupling MoS2,[314]
WSe2,[315,316] and GaSe[317] with optical microcavities leads to greatly enhanced PL
emitted in spectrally narrow and wavelength-tunable cavity modes and nanoscale,
optically-pumped lasing with a very low threshold. This technique should be further
developed to enhance device performance in broader wavelength regions for practical
applications.
6. Conclusions and Outlook
In summary, the large family of 2D layered semiconductors that can absorb light over
a broadband wavelengths from UV to NIR are ideal candidates for high-performance
photodetectors. Various types of photodetectors, including photoconductors,
phototransistors, and photodiodes, based on 2D layered semiconductors, have been
systematically investigated by many groups focusing on device design, performance,
and sensing mechanisms. More importantly, device performances have been
successfully optimized by various approaches, paving the way for highly sensitive
photodetectors with utility in many practical applications.
For photoconductors and phototransistors, devices based on 2D TMDs normally show
very high responsivities and gains but slow response times, making them suitable for
applications that only require high sensitivities. Devices based on other 2D layered
semiconductors, including group IIIA, group IVA, and group IVB metal chalcogenides,
generally show larger responsivities and faster response times than 2D MoS2-based
devices, rendering them promising candidates for sensitive photodetection in a
broadband wavelength region. Few-layer-BP-based detectors show large responsivities
in the UV-visible region and very fast responses at the telecommunication band,
indicating that this material is potentially useful for sensitive photodetection and optical
communication. Moreover, devices based on the hybrids of 2D layered semiconductors
have shown ultra-high responsivities up to 108 A W–1 and an extended detectable
wavelength range, providing an alternative means to realize high-performance devices.
Photodiodes based on 2D layered semiconductors have been realized with the following
device designs: (1) PN junctions based on 2D materials, including WSe2, MoSe2 and
few-layer BP, with properties controllable by local electrostatic gating; (2) Vertical or
lateral heterostructures with 2D layered semi-conductors sandwiched between
graphene layers or integrated with other 2D layered semiconductors by mechanical
stacking or direct synthesis; (3) Vertical or lateral homojunctions based on 2D layered
semiconductors with different doping regions;(4) van der Waal heterojunctions based
on the integration of a 2D layered semiconductor with other semiconductors such as
bulk Si; and (5) Schottky barrier diodes formed between 2D layered semiconductors
and metals. Although the responsivities of photodiodes are normally lower than those
of phototransistors, the response times of the former are much faster than those of the
latter devices. As evidenced by many examples, the 2D photodiodes have shown great
promise in photodetection and light harvesting, especially for flexible or
semitransparent systems. Moreover, manipulating photocarrier generation, separation,
and transport in the photodiodes has enabled us to better understand the underlying
working mechanisms and have pro-vided guidelines for device optimization.
In future work, there are many challenges left in this field. To realize high-performance
photodetectors, we need to have rational device designs and suitable materials for the
different parts of the device. We envision that 2D-material-based photodetectors can be
further developed in several directions. Firstly, novel 2D materials should be
synthesized and used in devices to meet some specific requirement for practical
applications. Secondly, novel approaches for chemical doping and surface treatment of
2D materials should be investigated, which is a feasible strategy for improving or
manipulating device performance. Thirdly, optical approaches, including plasmonic
technologies, optical waveguides, and microcavities, are very useful in enhancing the
light absorption of 2D materials and thus can be incorporated in the devices to increase
the sensitivities. For practical applications, the uniformity, stability, large-scale
production and cost of the devices are critical issues that must be tackled. More
importantly, 2D photodetectors should find some commercial applications at this stage
because their performances already exceed those of conventional semiconductor
devices; however, this has never been achieved until now. One feasible step towards
commercial applications is to integrate 2D materials with devices having the existing
pho-tonic and microelectronic platforms, such as CMOS technologies. This remains a
grand challenge to be solved in the future.
Acknowledgements
This work is financially supported by the Research Grants Council (RGC) of Hong
Kong, China (Project No. C5015-15G), and The Hong Kong Polytechnic University
(Project No. 1-BBA3, A-PL49).
Reference
[1] K. S. Novoselov, D. Jiang, F. Schedin, T. J. Booth, V. V Khotkevich, S. V
Morozov, A. K. Geim, Proc. Natl. Acad. Sci. 2005, 102, 10451.
[2] B. Radisavljevic, A. Radenovic, J. Brivio, V. Giacometti, A. Kis, Nat.
Nanotechnol. 2011, 6, 147.
[3] M. Xu, T. Liang, M. Shi, H. Chen, Chem. Rev. 2013, 113, 3766.
[4] S. Z. Butler, S. M. Hollen, L. Cao, Y. Cui, J. A. Gupta, H. R. Gutiérrez, T.
F. Heinz, S. S. Hong, J. Huang, A. F. Ismach, E. Johnston-Halperin, M. Kuno, V.
V. Plashnitsa, R. D. Robinson, R. S. Ruoff, S. Salahuddin, J. Shan, L. Shi, M. G.
Spencer, M. Terrones, W. Windl, J. E. Goldberger, ACS Nano 2013, 7, 2898.
[5] H. Li, J. Wu, Z. Yin, H. Zhang, Acc. Chem. Res. 2014, 47, 1067.
[6] M. Chhowalla, H. S. Shin, G. Eda, L.-J. Li, K. P. Loh, H. Zhang, Nat. Chem.
2013, 5, 263.
[7] G. Eda, H. Yamaguchi, D. Voiry, T. Fujita, M. Chen, M. Chhowalla, Nano
Lett. 2011, 11, 5111.
[8] J. N. Coleman, M. Lotya, A. O’Neill, S. D. Bergin, P. J. King, U. Khan, K.
Young, A. Gaucher, S. De, R. J. Smith, I. V Shvets, S. K. Arora, G. Stanton, H.-Y.
Kim, K. Lee, G. T. Kim, G. S. Duesberg, T. Hallam, J. J. Boland, J. J. Wang, J. F.
Donegan, J. C. Grunlan, G. Moriarty, A. Shmeliov, R. J. Nicholls, J. M. Perkins,
E. M. Grieveson, K. Theuwissen, D. W. McComb, P. D. Nellist, V. Nicolosi,
Science 2011, 331, 568.
[9] K. S. Novoselov, Science 2004, 306, 666.
[10] M. J. Allen, V. C. Tung, R. B. Kaner, Chem. Rev. 2010, 110, 132.
[11] K. S. Novoselov, V. I. Fal ko, L. Colombo, P. R. Gellert, M. G. Schwab,
K. Kim, Nature 2012, 490, 192.
[12] E. J. H. Lee, K. Balasubramanian, R. T. Weitz, M. Burghard, K. Kern, Nat.
Nanotechnol. 2008, 3, 486.
[13] K. F. Mak, L. Ju, F. Wang, T. F. Heinz, Solid State Commun. 2012, 152,
1341.
[14] F. Xia, H. Yan, P. Avouris, Proc. IEEE 2013, 101, 1717.
[15] Z. Sun, H. Chang, ACS Nano 2014, 8, 4133.
[16] F. H. L. Koppens, T. Mueller, P. Avouris, A. C. Ferrari, M. S. Vitiello, M.
Polini, Nat. Nanotechnol. 2014, 9, 780.
[17] J. Li, L. Niu, Z. Zheng, F. Yan, Adv. Mater. 2014, 26, 5239.
[18] F. Bonaccorso, Z. Sun, T. Hasan, A. C. Ferrari, Nat. Photonics 2010, 4, 611.
[19] F. Xia, T. Mueller, Y. Lin, A. Valdes-Garcia, P. Avouris, Nat. Nano-
technol. 2009, 4, 839.
[20] W. Bao, L. Jing, J. Velasco, Y. Lee, G. Liu, D. Tran, B. Standley, M. Aykol,
S. B. Cronin, D. Smirnov, M. Koshino, E. McCann, M. Bockrath, C. N. Lau, Nat.
Phys. 2011, 7, 948.
[21] A. Gupta, T. Sakthivel, S. Seal, Prog. Mater. Sci. 2015, 73, 44.
[22] Y. Shi, H. Li, L.-J. Li, Chem. Soc. Rev. 2015, 44, 2744.
[23] D. Jariwala, V. K. Sangwan, L. J. Lauhon, T. J. Marks, M. C. Hersam, ACS
Nano 2014, 8, 1102.
[24] X. Duan, C. Wang, A. Pan, R. Yu, X. Duan, Chem. Soc. Rev. 2015, 44,
8859.
[25] L. Li, Y. Yu, G. J. Ye, Q. Ge, X. Ou, H. Wu, D. Feng, X. H. Chen, Y. Zhang,
Nat. Nanotechnol. 2014, 9, 372.
[26] H. Liu, Y. Du, Y. Deng, P. D. Ye, Chem. Soc. Rev. 2015, 44, 2732.
[27] F. Xia, H. Wang, D. Xiao, M. Dubey, A. Ramasubramaniam, Nat. Photonics
2014, 8, 899.
[28] K. F. Mak, J. Shan, Nat. Photonics 2016, 10, 216.
[29] K. F. Mak, C. Lee, J. Hone, J. Shan, T. F. Heinz, Phys. Rev. Lett. 2010, 105,
136805.
[30] W. Jin, P.-C. Yeh, N. Zaki, D. Zhang, J. T. Sadowski, A. Al-Mahboob, A.
M. van der Zande, D. A. Chenet, J. I. Dadap, I. P. Herman, P. Sutter, J. Hone, R.
M. Osgood, Phys. Rev. Lett. 2013, 111, 106801.
[31] A. Ayari, E. Cobas, O. Ogundadegbe, M. S. Fuhrer, J. Appl. Phys. 2007,
101, 014507.
[32] A. Splendiani, L. Sun, Y. Zhang, T. Li, J. Kim, C.-Y. Chim, G. Galli, F.
Wang, Nano Lett. 2010, 10, 1271.
[33] H. Liu, A. T. Neal, Z. Zhu, Z. Luo, X. Xu, D. Tománek, P. D. Ye, ACS
Nano 2014, 8, 4033.
[34] V. Tran, R. Soklaski, Y. Liang, L. Yang, Phys. Rev. B 2014, 89, 235319.
[35] K. F. Mak, K. He, J. Shan, T. F. Heinz, Nat. Nanotechnol. 2012, 7, 494.
[36] K. F. Mak, K. He, C. Lee, G. H. Lee, J. Hone, T. F. Heinz, J. Shan, Nat.
Mater. 2012, 12, 207.
[37] M. Buscema, J. O. Island, D. J. Groenendijk, S. I. Blanter, G. A. Steele, H.
S. J. van der Zant, A. Castellanos-Gomez, Chem. Soc. Rev. 2015, 44, 3691.
[38] D. Akinwande, N. Petrone, J. Hone, Nat. Commun. 2014, 5, 5678.
[39] L. Britnell, R. M. Ribeiro, A. Eckmann, R. Jalil, B. D. Belle, Mishchenko,
Y.-J. Kim, R. V Gorbachev, T. Georgiou, S. V Morozov, A. N. Grigorenko, A. K.
Geim, C. Casiraghi, A. H. C. Neto, K. S. Novoselov, Science 2013, 340, 1311.
[40] W. S. Yun, S. W. Han, S. C. Hong, I. G. Kim, J. D. Lee, Phys. Rev. B 2012,
85, 033305.
[41] A. Castellanos-Gomez, M. Poot, G. A. Steele, H. S. J. van der Zant, N.
Agraït, G. Rubio-Bollinger, Adv. Mater. 2012, 24, 772.
[42] K. He, C. Poole, K. F. Mak, J. Shan, Nano Lett. 2013, 13, 2931.
[43] C. R. Zhu, G. Wang, B. L. Liu, X. Marie, X. F. Qiao, X. Zhang, X. X. Wu,
H. Fan, P. H. Tan, T. Amand, B. Urbaszek, Phys. Rev. B 2013, 88, 121301.
[44] A. K. Geim, I. V Grigorieva, Nature 2013, 499, 419.
[45] M.-Y. Li, C.-H. Chen, Y. Shi, L.-J. Li, Mater. Today 2015.
[46] S. Das, M. Kim, J. Lee, W. Choi, Crit. Rev. Solid State Mater. Sci. 2014,
39, 231.
[47] Y.-H. Lee, X.-Q. Zhang, W. Zhang, M.-T. Chang, C.-T. Lin, K.-D. Chang,
Y.-C. Yu, J. T.-W. Wang, C.-S. Chang, L.-J. Li, T.-W. Lin, Adv. Mater. 2012, 24,
2320.
[48] S. Das, J. Appenzeller, Appl. Phys. Lett. 2013, 103, 103501.
[49] N. R. Pradhan, D. Rhodes, Y. Xin, S. Memaran, L. Bhaskaran, M. Siddiq,
S. Hill, P. M. Ajayan, L. Balicas, ACS Nano 2014, 8, 7923.
[50] D. Kiriya, M. Tosun, P. Zhao, J. S. Kang, A. Javey, J. Am. Chem. Soc. 2014,
136, 7853.
[51] H. Fang, M. Tosun, G. Seol, T. C. Chang, K. Takei, J. Guo, A. Javey, Nano
Lett. 2013, 13, 1991.
[52] D. Xiang, C. Han, J. Wu, S. Zhong, Y. Liu, J. Lin, X.-A. Zhang, W. Ping
Hu, B. Özyilmaz, A. H. C. Neto, A. T. S. Wee, W. Chen, Nat. Commun. 2015, 6,
6485.
[53] S. Das, J. A. Robinson, M. Dubey, H. Terrones, M. Terrones, Annu. Rev.
Mater. Res. 2015, 45, 1.
[54] S. Tongay, J. Zhou, C. Ataca, K. Lo, T. S. Matthews, J. Li, J. C. Grossman,
J. Wu, Nano Lett. 2012, 12, 5576.
[55] T. Georgiou, R. Jalil, B. D. Belle, L. Britnell, R. V Gorbachev, S. V
Morozov, Y.-J. Kim, A. Gholinia, S. J. Haigh, O. Makarovsky, L. Eaves, L. A.
Ponomarenko, A. K. Geim, K. S. Novoselov, Mishchenko, Nat. Nanotechnol. 2012,
8, 100.
[56] A. M. Jones, H. Yu, N. J. Ghimire, S. Wu, G. Aivazian, J. S. Ross, Zhao, J.
Yan, D. G. Mandrus, D. Xiao, W. Yao, X. Xu, Nat. Nano- technol. 2013, 8, 634.
[57] Y. Li, Y. Rao, K. F. Mak, Y. You, S. Wang, C. R. Dean, T. F. Heinz, Nano
Lett. 2013, 13, 3329.
[58] G. Cunningham, M. Lotya, C. S. Cucinotta, S. Sanvito, S. D. Bergin, R.
Menzel, M. S. P. Shaffer, J. N. Coleman, ACS Nano 2012, 6, 3468.
[59] R. J. Smith, P. J. King, M. Lotya, C. Wirtz, U. Khan, S. De, A. O’Neill, G.
S. Duesberg, J. C. Grunlan, G. Moriarty, J. Chen, J. Wang, A. I. Minett, V. Nicolosi,
J. N. Coleman, Adv. Mater. 2011, 23, 3944.
[60] C. Zhi, Y. Bando, C. Tang, H. Kuwahara, D. Golberg, Adv. Mater. 2009,
21, 2889.
[61] Q. Tang, Z. Zhou, Prog. Mater. Sci. 2013, 58, 1244.
[62] J. Feng, L. Peng, C. Wu, X. Sun, S. Hu, C. Lin, J. Dai, J. Yang, Y. Xie, Adv.
Mater. 2012, 24, 1969.
[63] A. A. Green, M. C. Hersam, Nano Lett. 2009, 9, 4031.
[64] R. F. Frindt, A. S. Arrott, A. E. Curzon, B. Heinrich, S. R. Morrison, T. L.
Templeton, R. Divigalpitiya, M. A. Gee, P. Joensen, P. J. Schurer, J. L. LaCombe,
J. Appl. Phys. 1991, 70, 6224.
[65] H.-L. Tsai, J. Heising, J. L. Schindler, C. R. Kannewurf, M. G. Kanatzidis,
Chem. Mater. 1997, 9, 879.
[66] Z. Zeng, Z. Yin, X. Huang, H. Li, Q. He, G. Lu, F. Boey, H. Zhang, Angew.
Chemie Int. Ed. 2011, 50, 11093.
[67] Z. Zeng, T. Sun, J. Zhu, X. Huang, Z. Yin, G. Lu, Z. Fan, Q. Yan, H. H.
Hng, H. Zhang, Angew. Chemie Int. Ed. 2012, 51, 9052.
[68] J. Zhang, H. Yu, W. Chen, X. Tian, D. Liu, M. Cheng, G. Xie, W. Yang, R.
Yang, X. Bai, D. Shi, G. Zhang, ACS Nano 2014, 8, 6024.
[69] Y. H. Lee, L. Yu, H. Wang, W. Fang, X. Ling, Y. Shi, C. Te Lin, J. K.
Huang, M. T. Chang, C. S. Chang, M. Dresselhaus, T. Palacios, L. J. Li, J. Kong,
Nano Lett. 2013, 13, 1852.
[70] Y.-C. Lin, W. Zhang, J.-K. Huang, K.-K. Liu, Y.-H. Lee, C.-T. Liang, C.-
W. Chu, L.-J. Li, Nanoscale 2012, 4, 6637.
[71] Y. Zhan, Z. Liu, S. Najmaei, P. M. Ajayan, J. Lou, Small 2012, 8, 966.
[72] K. K. Liu, W. Zhang, Y. H. Lee, Y. C. Lin, M. T. Chang, C. Y. Su, C. S.
Chang, H. Li, Y. Shi, H. Zhang, C. S. Lai, L. J. Li, Nano Lett. 2012, 12, 1538.
[73] K. Kang, S. Xie, L. Huang, Y. Han, P. Y. Huang, K. F. Mak, C.-J. Kim, D.
Muller, J. Park, Nature 2015, 520, 656.
[74] R. H. Bube, Photoelectronic Properties of Semiconductors, Cam- bridge
University Press, Cambridge 1992.
[75] J. Watson, Opt. Laser Technol. 1992, 24, 177.
[76] G. Konstantatos, E. H. Sargent, Nat. Nanotechnol. 2010, 5, 391.
[77] G. Konstantatos, M. Badioli, L. Gaudreau, J. Osmond, M. Bernechea, F. P.
G. de Arquer, F. Gatti, F. H. L. Koppens, Nat. Nanotechnol. 2012, 7, 363.
[78] Z. Sun, Z. Liu, J. Li, G. Tai, S.-P. Lau, F. Yan, Adv. Mater. 2012, 24, 5878.
[79] S. M. Sze, K. K. Ng, Physics of Semiconductor Devices, 2nd Edition,
Wiley, New York 1981.
[80] X. Xu, N. M. Gabor, J. S. Alden, A. M. van der Zande, P. L. McEuen, Nano
Lett. 2010, 10, 562.
[81] J. C. W. Song, M. S. Rudner, C. M. Marcus, L. S. Levitov, Nano Lett. 2011,
11, 4688.
[82] R. W. Boyd, Radiometry and the Detection of Optical Radiation, John
Wiley & Sons, Hoboken, NJ 1983.
[83] P. L. Richards, J. Appl. Phys. 1994, 76, 1.
[84] M. Freitag, T. Low, F. Xia, P. Avouris, Nat. Photonics 2012, 7, 53.
[85] L. Vicarelli, M. S. Vitiello, D. Coquillat, A. Lombardo, A. C. Ferrari, W.
Knap, M. Polini, V. Pellegrini, A. Tredicucci, Nat. Mater. 2012, 11, 865.
[86] T. Low, A. S. Rodin, A. Carvalho, Y. Jiang, H. Wang, F. Xia, A. H. Castro
Neto, Phys. Rev. B 2014, 90, 075434.
[87] J. Liu, Photonic Devices, Cambridge University Press, Cambridge 2005.
[88] Z. Yin, H. H. Li, H. H. Li, L. Jiang, Y. Shi, Y. Sun, G. Lu, Q. Zhang, X.
Chen, H. Zhang, ACS Nano 2012, 6, 74.
[89] O. Lopez-Sanchez, D. Lembke, M. Kayci, A. Radenovic, A. Kis, Nat.
Nanotechnol. 2013, 8, 497.
[90] H. S. Lee, S.-W. Min, Y.-G. Chang, M. K. Park, T. Nam, H. Kim, J. H. Kim,
S. Ryu, S. Im, Nano Lett. 2012, 12, 3695.
[91] W. Choi, M. Y. Cho, A. Konar, J. H. Lee, G.-B. Cha, S. C. Hong, S. Kim,
J. Kim, D. Jena, J. Joo, S. Kim, Adv. Mater. 2012, 24, 5832.
[92] M. M. Furchi, D. K. Polyushkin, A. Pospischil, T. Mueller, Nano Lett. 2014,
14, 6165.
[93] D.-S. Tsai, D.-H. Lien, M.-L. Tsai, S.-H. Su, K.-M. Chen, J.-J. Ke, Y.-C.
Yu, L.-J. Li, J.-H. He, IEEE J. Sel. Top. Quantum Electron. 2014, 20, 30.
[94] M. F. Khan, M. W. Iqbal, M. Z. Iqbal, M. A. Shehzad, Y. Seo, J. Eom, ACS
Appl. Mater. Interfaces 2014, 6, 21645.
[95] W. Zhang, J.-K. Huang, C.-H. Chen, Y.-H. Chang, Y.-J. Cheng, L.-J. Li,
Adv. Mater. 2013, 25, 3456.
[96] F. Schedin, A. K. Geim, S. V. Morozov, E. W. Hill, P. Blake, M. I.
Katsnelson, K. S. Novoselov, Nat. Mater. 2007, 6, 652.
[97] S. Tongay, J. Zhou, C. Ataca, J. Liu, J. S. Kang, T. S. Matthews, L. You, J.
Li, J. C. Grossman, J. Wu, Nano Lett. 2013, 13, 2831.
[98] N. Perea-López, Z. Lin, N. R. Pradhan, A. Iñiguez-Rábago, A. Laura Elías,
A. McCreary, J. Lou, P. M. Ajayan, H. Terrones, L. Balicas, M. Terrones, 2D
Mater. 2014, 1, 011004.
[99] D.-S. Tsai, K.-K. Liu, D.-H. Lien, M.-L. Tsai, C.-F. Kang, C.-A. Lin, L.-J.
Li, J.-H. He, ACS Nano 2013, 7, 3905.
[100] J. Li, M. M. Naiini, S. Vaziri, M. C. Lemme, M. Östling, Adv. Funct. Mater.
2014, 24, 6524.
[101] G. Cunningham, U. Khan, C. Backes, D. Hanlon, D. McCloskey, J. F.
Donegan, J. N. Coleman, J. Mater. Chem. C 2013, 1, 6899.
[102] Z. P. Ling, R. Yang, J. W. Chai, S. J. Wang, W. S. Leong, Y. Tong, D. Lei,
Q. Zhou, X. Gong, D. Z. Chi, K.-W. Ang, Opt. Express 2015, 23, 13580.
[103] Y. Lee, J. Yang, D. Lee, Y.-H. Kim, J.-H. Park, H. Kim, J. H. Cho,
Nanoscale 2016, 8, 9193.
[104] J. Lu, J. H. Lu, H. Liu, B. Liu, K. X. Chan, J. Lin, W. Chen, K. P. Loh, C.
H. Sow, ACS Nano 2014, 8, 6334.
[105] J. Kwon, Y. K. Hong, G. Han, I. Omkaram, W. Choi, S. Kim, Y. Yoon, Adv.
Mater. 2015, 27, 2224.
[106] D. Kufer, G. Konstantatos, Nano Lett. 2015, 15, 7307.
[107] X. Wang, P. Wang, J. Wang, W. Hu, X. Zhou, N. Guo, H. Huang, S. Sun,
H. Shen, T. Lin, M. Tang, L. Liao, A. Jiang, J. Sun, X. Meng, X. Chen, W. Lu, J.
Chu, Adv. Mater. 2015, 27, 6575.
[108] H. S. Lee, S. S. Baik, K. Lee, S. Min, P. J. Jeon, J. S. Kim, K. Choi, H. J.
Choi, J. H. Kim, S. Im, ACS Nano 2015, 9, 8312.
[109] Y. Pang, F. Xue, L. Wang, J. Chen, J. Luo, T. Jiang, C. Zhang, Z. L. Wang,
Adv. Sci. 2016, 3, 1500419.
[110] J. D. Lin, C. Han, F. Wang, R. Wang, D. Xiang, S. Qin, X. A. Zhang, L.
Wang, H. Zhang, A. T. S. Wee, W. Chen, ACS Nano 2014, 8, 5323.
[111] D. H. Kang, M. S. Kim, J. Shim, J. Jeon, H. Y. Park, W. S. Jung, H. Y. Yu,
C. H. Pang, S. Lee, J. H. Park, Adv. Funct. Mater. 2015, 25, 4219.
[112] A. Ramasubramaniam, D. Naveh, E. Towe, Phys. Rev. B 2011, 84, 205325.
[113] N. Zibouche, P. Philipsen, A. Kuc, T. Heine, Phys. Rev. B 2014, 90, 125440.
[114] H. Wang, C. Zhang, W. Chan, S. Tiwari, F. Rana, Nat. Commun. 2015, 6,
8831.
[115] M. Buscema, M. Barkelid, V. Zwiller, H. S. J. van der Zant, G. A. Steele,
A. Castellanos-Gomez, Nano Lett. 2013, 13, 358.
[116] Y. Zhang, H. Li, L. Wang, H. Wang, X. Xie, S.-L. Zhang, R. Liu, Z.-J. Qiu,
Sci. Rep. 2015, 5, 7938.
[117] Y. Zhang, T.-R. Chang, B. Zhou, Y.-T. Cui, H. Yan, Z. Liu, F. Schmitt, J.
Lee, R. Moore, Y. Chen, H. Lin, H.-T. Jeng, S.-K. Mo, Z. Hussain, A. Bansil, Z.-
X. Shen, Nat. Nanotechnol. 2013, 9, 111.
[118] J. S. Ross, S. Wu, H. Yu, N. J. Ghimire, A. M. Jones, G. Aivazian, J. Yan,
D. G. Mandrus, D. Xiao, W. Yao, X. Xu, Nat. Commun. 2013, 4, 1474.
[119] S. Larentis, B. Fallahazad, E. Tutuc, Appl. Phys. Lett. 2012, 101, 223104.
[120] A. Abderrahmane, P. J. Ko, T. V Thu, S. Ishizawa, T. Takamura, A. Sandhu,
Nanotechnology 2014, 25, 365202.
[121] X. Lu, M. I. B. Utama, J. Lin, X. Gong, J. Zhang, Y. Zhao, S. T. Pantelides,
J. Wang, Z. Dong, Z. Liu, W. Zhou, Q. Xiong, Nano Lett. 2014, 14, 2419.
[122] X. Wang, Y. Gong, G. Shi, W. L. Chow, K. Keyshar, G. Ye, R. Vajtai, J.
Lou, Z. Liu, E. Ringe, B. K. Tay, P. M. Ajayan, ACS Nano 2014, 8, 5125.
[123] G. W. Shim, K. Yoo, S.-B. Seo, J. Shin, D. Y. Jung, I.-S. Kang, C. W. Ahn,
B. J. Cho, S.-Y. Choi, ACS Nano 2014, 8, 6655.
[124] Y.-H. Chang, W. Zhang, Y. Zhu, Y. Han, J. Pu, J.-K. Chang, W.-T. Hsu, J.-
K. Huang, C.-L. Hsu, M.-H. Chiu, T. Takenobu, H. Li, C.-I. Wu, W.-H. Chang, A.
T. S. Wee, L.-J. Li, ACS Nano 2014, 8, 8582.
[125] J. Xia, X. Huang, L.-Z. Liu, M. Wang, L. Wang, B. Huang, D.-D. Zhu, J.-J.
Li, C.-Z. Gu, X.-M. Meng, Nanoscale 2014, 6, 8949.
[126] C. Jung, S. M. Kim, H. Moon, G. Han, J. Kwon, Y. K. Hong, I. Omkaram,
Y. Yoon, S. Kim, J. Park, Sci. Rep. 2015, 5, 15313.
[127] W. Sik Hwang, M. Remskar, R. Yan, V. Protasenko, K. Tahy, S. Doo Chae,
P. Zhao, A. Konar, H. (Grace) Xing, A. Seabaugh, D. Jena, Appl. Phys. Lett. 2012,
101, 013107.
[128] S. Hwan Lee, D. Lee, W. Sik Hwang, E. Hwang, D. Jena, W. Jong Yoo,
Appl. Phys. Lett. 2014, 104, 193113.
[129] N. Perea-López, A. L. Elías, A. Berkdemir, A. Castro-Beltran, H. R.
Gutiérrez, S. Feng, R. Lv, T. Hayashi, F. López-Urías, S. Ghosh, B. Muchharla, S.
Talapatra, H. Terrones, M. Terrones, Adv. Funct. Mater. 2013, 23, 5511.
[130] N. Huo, S. Yang, Z. Wei, S.-S. Li, J.-B. Xia, J. Li, Sci. Rep. 2014, 4, 5209.
[131] C. Lan, C. Li, Y. Yin, Y. Liu, Nanoscale 2015, 7, 5974.
[132] J. D. Yao, Z. Q. Zheng, J. M. Shao, G. W. Yang, Nanoscale 2015, 7, 14974.
[133] L. Zeng, L. Tao, C. Tang, B. Zhou, H. Long, Y. Chai, S. P. Lau, Y. H. Tsang,
Sci. Rep. 2016, 6, 20343.
[134] W. Zhang, M.-H. Chiu, C.-H. Chen, W. Chen, L.-J. Li, A. T. S. Wee, ACS
Nano 2014, 8, 8653.
[135] N. R. Pradhan, J. Ludwig, Z. Lu, D. Rhodes, M. M. Bishop, K.
Thirunavukkuarasu, S. A. McGill, D. Smirnov, L. Balicas, ACS Appl. Mater.
Interfaces 2015, 7, 12080.
[136] J. Chen, B. Liu, Y. Liu, W. Tang, C. T. Nai, L. Li, J. Zheng, L. Gao, Y.
Zheng, H. S. Shin, H. Y. Jeong, K. P. Loh, Adv. Mater. 2015, 27, 6722.
[137] S.-H. Jo, D.-H. Kang, J. Shim, J. Jeon, M. H. Jeon, G. Yoo, J. Kim, J. Lee,
G. Y. Yeom, S. Lee, H.-Y. Yu, C. Choi, J.-H. Park, Adv. Mater. 2016, 28, 4824.
[138] Y.-F. Lin, Y. Xu, S.-T. Wang, S.-L. Li, M. Yamamoto, A. Aparecido-
Ferreira, W. Li, H. Sun, S. Nakaharai, W.-B. Jian, K. Ueno, K. Tsukagoshi, Adv.
Mater. 2014, 26, 3263.
[139] C. Ruppert, O. B. Aslan, T. F. Heinz, Nano Lett. 2014, 14, 6231.
[140] N. R. Pradhan, D. Rhodes, S. Feng, Y. Xin, S. Memaran, B.-H. Moon, H.
Terrones, M. Terrones, L. Balicas, ACS Nano 2014, 8, 5911.
[141] L. Yin, X. Zhan, K. Xu, F. Wang, Z. Wang, Y. Huang, Q. Wang, C. Jiang,
J. He, Appl. Phys. Lett. 2016, 108, 043503.
[142] K. Xu, Z. Wang, F. Wang, Y. Huang, F. Wang, L. Yin, C. Jiang, J. He, Adv.
Mater. 2015, 27, 7881.
[143] J. A. Wilson, A. D. Yoffe, Adv. Phys. 1969, 18, 193.
[144] S. Yang, S. Tongay, Y. Li, Q. Yue, J.-B. Xia, S.-S. Li, J. Li, S.-H. Wei,
Nanoscale 2014, 6, 7226.
[145] S. Yang, S. Tongay, Q. Yue, Y. Li, B. Li, F. Lu, Sci. Rep. 2014, 4, 5442.
[146] E. Zhang, Y. Jin, X. Yuan, W. Wang, C. Zhang, L. Tang, S. Liu, P. Zhou,
W. Hu, F. Xiu, Adv. Funct. Mater. 2015, 25, 4076.
[147] F. Liu, S. Zheng, X. He, A. Chaturvedi, J. He, W. L. Chow, T. R. Mion, X.
Wang, J. Zhou, Q. Fu, H. J. Fan, B. K. Tay, L. Song, R.-H. He, C. Kloc, P. M.
Ajayan, Z. Liu, Adv. Funct. Mater. 2016, 26, 1169.
[148] M. Hafeez, L. Gan, H. Li, Y. Ma, T. Zhai, Adv. Funct. Mater. 2016, 26,
4551.
[149] E. Liu, M. Long, J. Zeng, W. Luo, Y. Wang, Y. Pan, W. Zhou, B. Wang,
W. Hu, Z. Ni, Y. You, X. Zhang, S. Qin, Y. Shi, K. Watanabe, T. Taniguchi, H.
Yuan, H. Y. Hwang, Y. Cui, F. Miao, D. Xing, Adv. Funct. Mater. 2016, 26, 1938.
[150] S. Jo, H.-Y. Park, D. Kang, J. Shim, J. Jeon, S. Choi, M. Kim, Y. Park, J.
Lee, Y. J. Song, S. Lee, J. Park, Adv. Mater. 2016, 28, 6711.
[151] J. Shim, A. Oh, D.-H. Kang, S. Oh, S. K. Jang, J. Jeon, M. H. Jeon, M. Kim,
C. Choi, J. Lee, S. Lee, G. Y. Yeom, Y. J. Song, J.-H. Park, Adv. Mater. 2016, 28,
6985.
[152] P. Hu, L. Wang, M. Yoon, J. Zhang, W. Feng, X. Wang, Z. Wen, J. C.
Idrobo, Y. Miyamoto, D. B. Geohegan, K. Xiao, Nano Lett. 2013, 13, 1649.
[153] P. Hu, Z. Wen, L. Wang, P. Tan, K. Xiao, ACS Nano 2012, 6, 5988.
[154] P. Hu, J. Zhang, M. Yoon, X.-F. Qiao, X. Zhang, W. Feng, P. Tan, W.
Zheng, J. Liu, X. Wang, J. C. Idrobo, D. B. Geohegan, K. Xiao, Nano Res. 2014,
7, 694.
[155] F. Liu, H. Shimotani, H. Shang, T. Kanagasekaran, V. Zólyomi, N.
Drummond, V. I. Fal’ko, K. Tanigaki, ACS Nano 2014, 8, 752.
[156] S. Yang, Y. Li, X. Wang, N. Huo, J.-B. Xia, S.-S. Li, J. Li, Nanoscale 2014,
6, 2582.
[157] Z. Wang, K. Xu, Y. Li, X. Zhan, M. Safdar, Q. Wang, F. Wang, J. He, ACS
Nano 2014, 8, 4859.
[158] Y. Cao, K. Cai, P. Hu, L. Zhao, T. Yan, W. Luo, X. Zhang, X. Wu, K. Wang,
H. Zheng, Sci. Rep. 2015, 5, 8130.
[159] S. Lei, L. Ge, Z. Liu, S. Najmaei, G. Shi, G. You, J. Lou, R. Vajtai, P. M.
Ajayan, Nano Lett. 2013, 13, 2777.
[160] M. Mahjouri-Samani, R. Gresback, M. Tian, K. Wang, A. A.
Puretzky, C. M. Rouleau, G. Eres, I. N. Ivanov, K. Xiao, M. A. McGuire, G.
Duscher, D. B. Geohegan, Adv. Funct. Mater. 2014, 24, 6365.
[161] X. Li, M.-W. Lin, A. A. Puretzky, J. C. Idrobo, C. Ma, M. Chi, M. Yoon, C.
M. Rouleau, I. I. Kravchenko, D. B. Geohegan, K. Xiao, Sci. Rep. 2014, 4, 5497.
[162] Z. Wang, M. Safdar, M. Mirza, K. Xu, Q. Wang, Y. Huang, F. Wang, X.
Zhan, J. He, Nanoscale 2015, 7, 7252.
[163] M. Lin, D. Wu, Y. Zhou, W. Huang, W. Jiang, W. Zheng, S. Zhao, C. Jin,
Y. Guo, H. Peng, Z. Liu, J. Am. Chem. Soc. 2013, 135, 13274.
[164] R. B. Jacobs-Gedrim, M. Shanmugam, N. Jain, C. A. Durcan, M. T.
Murphy, T. M. Murray, R. J. Matyi, R. L. Moore, B. Yu, ACS Nano 2014, 8, 514.
[165] J. Zhou, Q. Zeng, D. Lv, L. Sun, L. Niu, W. Fu, F. Liu, Z. Shen, C. Jin, Z.
Liu, Nano Lett. 2015, 15, 6400.
[166] J. O. Island, S. I. Blanter, M. Buscema, H. S. J. van der Zant, A. Castellanos-
Gomez, Nano Lett. 2015, 15, 7853.
[167] W. Zheng, T. Xie, Y. Zhou, Y. L. Chen, W. Jiang, S. Zhao, J. Wu, Y. Jing,
Y. Wu, G. Chen, Y. Guo, J. Yin, S. Huang, H. Q. Xu, Z. Liu, H. Peng, Nat.
Commun. 2015, 6, 6972.
[168] S. Lei, L. Ge, S. Najmaei, A. George, R. Kappera, J. Lou, M. Chhowalla, H.
Yamaguchi, G. Gupta, R. Vajtai, A. D. Mohite, P. M. Ajayan, ACS Nano 2014, 8,
1263.
[169] S. R. Tamalampudi, Y.-Y. Lu, U. R. Kumar, R. Sankar, C.-D. Liao, B. K.
Moorthy, C.-H. Cheng, F. C. Chou, Y.-T. Chen, Nano Lett. 2014, 14, 2800.
[170] W. Feng, J.-B. Wu, X. Li, W. Zheng, X. Zhou, K. Xiao, W. Cao, B. Yang,
J.-C. Idrobo, L. Basile, W. Tian, P. Tan, P. Hu, J. Mater. Chem. C 2015, 3, 7022.
[171] W. Luo, Y. Cao, P. Hu, K. Cai, Q. Feng, F. Yan, T. Yan, X. Zhang, K.
Wang, Adv. Opt. Mater. 2015, 3, 1418.
[172] S. Lei, F. Wen, L. Ge, S. Najmaei, A. George, Y. Gong, W. Gao, Z. Jin, B.
Li, J. Lou, J. Kono, R. Vajtai, P. Ajayan, N. J. Halas, Nano Lett. 2015, 15, 3048.
[173] J. Jasinski, W. Swider, J. Washburn, Z. Liliental-Weber, A. Chaiken, K.
Nauka, G. A. Gibson, C. C. Yang, Appl. Phys. Lett. 2002, 81, 4356.
[174] T. Zhai, X. Fang, M. Liao, X. Xu, L. Li, B. Liu, Y. Koide, Y. Ma, J. Yao,
Y. Bando, D. Golberg, ACS Nano 2010, 4, 1596.
[175] K. Lai, H. Peng, W. Kundhikanjana, D. T. Schoen, C. Xie, S. Meister, Y.
Cui, M. A. Kelly, Z.-X. Shen, Nano Lett. 2009, 9, 1265.
[176] G. W. Mudd, S. A. Svatek, T. Ren, A. Patanè, O. Makarovsky, L. Eaves, P.
H. Beton, Z. D. Kovalyuk, G. V. Lashkarev, Z. R. Kudrynskyi, A. I. Dmitriev, Adv.
Mater. 2013, 25, 5714.
[177] Y. Huang, E. Sutter, J. T. Sadowski, M. Cotlet, O. L. A. Monti, D. A. Racke,
M. R. Neupane, D. Wickramaratne, R. K. Lake, B. A. Parkinson, P. Sutter, ACS
Nano 2014, 8, 10743.
[178] G. Su, V. G. Hadjiev, P. E. Loya, J. Zhang, S. Lei, S. Maharjan, P. Dong, P.
M. Ajayan, J. Lou, H. Peng, Nano Lett. 2015, 15, 506.
[179] Y. Huang, H.-X. Deng, K. Xu, Z.-X. Wang, Q.-S. Wang, F.-M. Wang, F.
Wang, X.-Y. Zhan, S.-S. Li, J.-W. Luo, J. He, Nanoscale 2015, 7, 14093.
[180] D. Yang, B. Li, C. Hu, H. Deng, D. Dong, X. Yang, K. Qiao, S. Yuan, H.
Song, Adv. Opt. Mater. 2016, 4, 419.
[181] J. Xia, D. Zhu, L. Wang, B. Huang, X. Huang, X.-M. Meng, Adv. Funct.
Mater. 2015, 25, 4255.
[182] Y. Tao, X. Wu, W. Wang, J. Wang, J. Mater. Chem. C 2015, 3, 1347.
[183] X. Zhou, L. Gan, W. Tian, Q. Zhang, S. Jin, H. Li, Y. Bando, D. Golberg,
T. Zhai, Adv. Mater. 2015, 27, 8035.
[184] Y. Huang, K. Xu, Z. Wang, T. A. Shifa, Q. Wang, F. Wang, C. Jiang, J. He,
Nanoscale 2015, 7, 17375.
[185] P. Yu, X. Yu, W. Lu, H. Lin, L. Sun, K. Du, F. Liu, W. Fu, Q. Zeng, Z.
Shen, C. Jin, Q. J. Wang, Z. Liu, Adv. Funct. Mater. 2016, 26, 137.
[186] D.-J. Xue, J. Tan, J.-S. Hu, W. Hu, Y.-G. Guo, L.-J. Wan, Adv. Mater. 2012,
24, 4528.
[187] R. K. Ulaganathan, Y.-Y. Lu, C.-J. Kuo, S. R. Tamalampudi, R. Sankar, K.
M. Boopathi, A. Anand, K. Yadav, R. J. Mathew, C.-R. Liu, F. C. Chou, Y.-T.
Chen, Nanoscale 2016, 8, 2284.
[188] R. Schlaf, N. Armstrong, B. Parkinson, C. Pettenkofer, W. Jaegermann,
Surf. Sci. 1997, 385, 1.
[189] Y. Sun, H. Cheng, S. Gao, Z. Sun, Q. Liu, Q. Liu, F. Lei, T. Yao, J. He, S.
Wei, Y. Xie, Angew. Chemie Int. Ed. 2012, 51, 8727.
[190] S. G. Hickey, C. Waurisch, B. Rellinghaus, A. Eychmüller, J. Am. Chem.
Soc. 2008, 130, 14978.
[191] D. D. VaughnII, R. J. Patel, M. A. Hickner, R. E. Schaak, J. Am. Chem.
Soc. 2010, 132, 15170.
[192] L. Li, Z. Chen, Y. Hu, X. Wang, T. Zhang, W. Chen, Q. Wang, J. Am.
Chem. Soc. 2013, 135, 1213.
[193] J. Kang, H. Sahin, H. D. Ozaydin, R. T. Senger, F. M. Peeters, Phys. Rev.
B 2015, 92, 075413.
[194] W. Schairer, M. W. Shafer, Phys. Status Solidi 1973, 17, 181.
[195] D. Pacilé, M. Papagno, M. Lavagnini, H. Berger, L. Degiorgi, M. Grioni,
Phys. Rev. B 2007, 76, 155406.
[196] J. O. Island, M. Buscema, M. Barawi, J. M. Clamagirand, J. R. Ares, C.
Sánchez, I. J. Ferrer, G. A. Steele, H. S. J. van der Zant, A. Castellanos-Gomez,
Adv. Opt. Mater. 2014, 2, 641.
[197] Y.-R. Tao, X.-C. Wu, W.-W. Xiong, Small 2014, 10, 4905.
[198] Y.-R. Tao, J.-J. Wu, X.-C. Wu, Nanoscale 2015, 7, 14292.
[199] W.-W. Xiong, J.-Q. Chen, X.-C. Wu, J.-J. Zhu, J. Mater. Chem. C 2014, 2,
7392.
[200] W.-W. Xiong, J.-Q. Chen, X.-C. Wu, J.-J. Zhu, J. Mater. Chem. C 2015, 3,
1929.
[201] L. M. Xie, Nanoscale 2015, 7, 18392.
[202] C. Tan, P. Yu, Y. Hu, J. Chen, Y. Huang, Y. Cai, Z. Luo, B. Li, Q. Lu, L.
Wang, Z. Liu, H. Zhang, J. Am. Chem. Soc. 2015, 137, 10430.
[203] W. Feng, W. Zheng, P. Hu, Phys. Chem. Chem. Phys. 2014, 16, 19340.
[204] Q. Wang, K. Xu, Z. Wang, F. Wang, Y. Huang, M. Safdar, X. Zhan, F.
Wang, Z. Cheng, J. He, Nano Lett. 2015, 15, 1183.
[205] V. Klee, E. Preciado, D. Barroso, A. E. Nguyen, C. Lee, K. J. Erickson, M.
Triplett, B. Davis, I. H. Lu, S. Bobek, J. McKinley, J. P. Martinez, J. Mann, A. A.
Talin, L. Bartels, F. Léonard, Nano Lett. 2015, 15, 2612.
[206] N. Kohara, S. Nishiawaki, T. Negami, T. Wada, Jpn. J. Appl. Phys. 2000,
39, 6316.
[207] S. Lei, A. Sobhani, F. Wen, A. George, Q. Wang, Y. Huang, P. Dong, B. Li,
S. Najmaei, J. Bellah, G. Gupta, A. D. Mohite, L. Ge, J. Lou, N. J. Halas, R. Vajtai,
P. Ajayan, Adv. Mater. 2014, 26, 7666.
[208] P. Perumal, R. K. Ulaganathan, R. Sankar, Y.-M. Liao, T.-M. Sun, M.-W.
Chu, F. C. Chou, Y.-T. Chen, M.-H. Shih, Y.-F. Chen, Adv. Funct. Mater. 2016,
26, 3630.
[209] Y. Akahama, S. Endo, S. Narita, J. Phys. Soc. Japan 1983, 52, 2148.
[210] S. Narita, Y. Akahama, Y. Tsukiyama, K. Muro, S. Mori, S. Endo, M.
Taniguchi, M. Seki, S. Suga, A. Mikuni, H. Kanzaki, Phys. B C 1983, 117–118,
422.
[211] Y. Maruyama, S. Suzuki, K. Kobayashi, S. Tanuma, Phys. B+C 1981, 105,
99.
[212] F. Xia, H. Wang, Y. Jia, Nat. Commun. 2014, 5, 4458.
[213] A. S. Rodin, A. Carvalho, A. H. Castro Neto, Phys. Rev. Lett. 2014, 112,
176801.
[214] J. Qiao, X. Kong, Z.-X. Hu, F. Yang, W. Ji, Nat. Commun. 2014, 5, 4475.
[215] X. Wang, A. M. Jones, K. L. Seyler, V. Tran, Y. Jia, H. Zhao, H. Wang, L.
Yang, X. Xu, F. Xia, Nat. Nanotechnol. 2015, 10, 517.
[216] M. Buscema, D. J. Groenendijk, S. I. Blanter, G. A. Steele, H. S. J. van der
Zant, A. Castellanos-Gomez, Nano Lett. 2014, 14, 3347.
[217] T. Hong, B. Chamlagain, W. Lin, H.-J. Chuang, M. Pan, Z. Zhou, Y.-Q. Xu,
Nanoscale 2014, 6, 8978.
[218] J. Wu, G. K. W. Koon, D. Xiang, C. Han, C. T. Toh, E. S. Kulkarni,
Verzhbitskiy, A. Carvalho, A. S. Rodin, S. P. Koenig, G. Eda, Chen, A. H. C. Neto,
B. Özyilmaz, ACS Nano 2015, 9, 8070.
[219] M. Huang, M. Wang, C. Chen, Z. Ma, X. Li, J. Han, Y. Wu, Adv. Mater.
2016, 28, 3481.
[220] M. Engel, M. Steiner, P. Avouris, Nano Lett. 2014, 14, 6414.
[221] L. Viti, J. Hu, D. Coquillat, W. Knap, A. Tredicucci, A. Politano, M. S.
Vitiello, Adv. Mater. 2015, 27, 5567.
[222] K. Roy, M. Padmanabhan, S. Goswami, T. P. Sai, G. Ramalingam, S.
Raghavan, A. Ghosh, Nat. Nanotechnol. 2013, 8, 826.
[223] W. Zhang, C.-P. Chuu, J.-K. Huang, C.-H. Chen, M.-L. Tsai, Y.-H. Chang,
C.-T. Liang, Y.-Z. Chen, Y.-L. Chueh, J.-H. He, M.-Y. Chou, L.-J. Li, Sci. Rep.
2014, 4, 3826.
[224] H. Xu, J. Wu, Q. Feng, N. Mao, C. Wang, J. Zhang, Small 2014, 10, 2300.
[225] X. Li, J. Wu, N. Mao, J. Zhang, Z. Lei, Z. Liu, H. Xu, Carbon 2015, 92, 126.
[226] D. Ma, J. Shi, Q. Ji, K. Chen, J. Yin, Y. Lin, Y. Zhang, M. Liu, Q. Feng, X.
Song, X. Guo, J. Zhang, Y. Zhang, Z. Liu, Nano Res. 2015, 8, 3662.
[227] J. Yang, H. Kwak, Y. Lee, Y.-S. Kang, M.-H. Cho, J. H. Cho, Y.-H. Kim,
S.-J. Jeong, S. Park, H.-J. Lee, H. Kim, ACS Appl. Mater. Interfaces 2016, 8, 8576.
[228] H. Qiao, J. Yuan, Z. Xu, C. Chen, S. Lin, Y. Wang, J. Song, Y. Liu, Q.
Khan, H. Y. Hoh, C.-X. Pan, S. Li, Q. Bao, ACS Nano 2015, 9, 1886.
[229] Z. Chen, J. Biscaras, A. Shukla, Nanoscale 2015, 7, 5981.
[230] D. Kufer, I. Nikitskiy, T. Lasanta, G. Navickaite, F. H. L. Koppens, G.
Konstantatos, Adv. Mater. 2015, 27, 176.
[231] J. Schornbaum, B. Winter, S. P. Schießl, F. Gannott, G. Katsukis, D. M.
Guldi, E. Spiecker, J. Zaumseil, Adv. Funct. Mater. 2014, 24, 5798.
[232] C. Chen, H. Qiao, S. Lin, C. Man Luk, Y. Liu, Z. Xu, J. Song, Y. Xue, D.
Li, J. Yuan, W. Yu, C. Pan, S. Ping Lau, Q. Bao, Sci. Rep. 2015, 5, 11830.
[233] Y. Huang, X. Zhan, K. Xu, L. Yin, Z. Cheng, C. Jiang, Z. Wang, J. He,
Appl. Phys. Lett. 2016, 108, 013101.
[234] S. H. Yu, Y. Lee, S. K. Jang, J. Kang, J. Jeon, C. Lee, J. Y. Lee, H. Kim, E.
Hwang, S. Lee, J. H. Cho, ACS Nano 2014, 8, 8285.
[235] E. H. Cho, W. G. Song, C. J. Park, J. Kim, S. Kim, J. Joo, Nano Res. 2015,
8, 790.
[236] C. Ma, Y. Shi, W. Hu, M. H. Chiu, Z. Liu, A. Bera, F. Li, H. Wang, L. J.
Li, T. Wu,. Adv. Mater. 2016, 28, 3683.
[237] D. Kang, S. R. Pae, J. Shim, G. Yoo, J. Jeon, J. W. Leem, J. S. Yu, S. Lee,
B. Shin, J.-H. Park, Adv. Mater. 2016, DOI: 10.1002/ adma.201600992.
[238] A. Pospischil, M. M. Furchi, T. Mueller, Nat. Nanotechnol. 2014, 9, 257.
[239] B. W. H. Baugher, H. O. H. Churchill, Y. Yang, P. Jarillo-Herrero, Nat.
Nanotechnol. 2014, 9, 262.
[240] J. S. Ross, P. Klement, A. M. Jones, N. J. Ghimire, J. Yan, D. G. Mandrus,
T. Taniguchi, K. Watanabe, K. Kitamura, W. Yao, D. H. Cobden, X. Xu, Nat.
Nanotechnol. 2014, 9, 268.
[241] D. J. Groenendijk, M. Buscema, G. A. Steele, S. Michaelis de Vas-
concellos, R. Bratschitsch, H. S. J. van der Zant, A. Castellanos- Gomez, Nano Lett.
2014, 14, 5846.
[242] M. Buscema, D. J. Groenendijk, G. a Steele, H. S. J. van der Zant, A.
Castellanos-Gomez, Nat. Commun. 2014, 5, 4651.
[243] S. Memaran, N. R. Pradhan, Z. Lu, D. Rhodes, J. Ludwig, Q. Zhou, O.
Ogunsolu, P. M. Ajayan, D. Smirnov, A. I. Fernández- Domínguez, F. J. García-
Vidal, L. Balicas, Nano Lett. 2015, 15, 7532.
[244] C. R. Dean, A. F. Young, I. Meric, C. Lee, L. Wang, S. Sorgenfrei, K.
Watanabe, T. Taniguchi, P. Kim, K. L. Shepard, J. Hone, Nat. Nanotechnol. 2010,
5, 722.
[245] F. Bonaccorso, A. Lombardo, T. Hasan, Z. Sun, L. Colombo, A. C. Ferrari,
Mater. Today 2012, 15, 564.
[246] W. J. Yu, Y. Liu, H. Zhou, A. Yin, Z. Li, Y. Huang, X. Duan, Nat.
Nanotechnol. 2013, 8, 952.
[247] M. Massicotte, P. Schmidt, F. Vialla, K. G. Schädler, A. Reserbat- Plantey,
K. Watanabe, T. Taniguchi, K. J. Tielrooij, F. H. L. Koppens, Nat. Nanotechnol.
2015, 11, 42.
[248] G. W. Mudd, S. A. Svatek, L. Hague, O. Makarovsky, Z. R. Kudrynskyi, C.
J. Mellor, P. H. Beton, L. Eaves, K. S. Novoselov, Z. D. Kovalyuk, E. E. Vdovin,
A. J. Marsden, N. R. Wilson, A. Patanè, Adv. Mater. 2015, 27, 3760.
[249] H. Fang, C. Battaglia, C. Carraro, S. Nemsak, B. Ozdol, J. S. Kang, H. A.
Bechtel, S. B. Desai, F. Kronast, A. A. Unal, G. Conti, C. Conlon, G. K. Palsson,
M. C. Martin, A. M. Minor, C. S. Fadley, E. Yablonovitch, R. Maboudian, A.
Javey, Proc. Natl. Acad. Sci. 2014, 111, 6198.
[250] X. Hong, J. Kim, S.-F. Shi, Y. Zhang, C. Jin, Y. Sun, S. Tongay, J. Wu, Y.
Zhang, F. Wang, Nat. Nanotechnol. 2014, 9, 682.
[251] F. Ceballos, M. Z. Bellus, H.-Y. Chiu, H. Zhao, ACS Nano 2014, 8, 12717.
[252] R. Long, O. V. Prezhdo, Nano Lett. 2016, 16, 1996.
[253] C.-H. Lee, G. Lee, A. M. van der Zande, W. Chen, Y. Li, M. Han, Cui, G.
Arefe, C. Nuckolls, T. F. Heinz, J. Guo, J. Hone, P. Kim, Nat. Nanotechnol. 2014,
9, 676.
[254] M. M. Furchi, A. Pospischil, F. Libisch, J. Burgdörfer, T. Mueller, Nano
Lett. 2014, 14, 4785.
[255] R. Cheng, D. Li, H. Zhou, C. Wang, A. Yin, S. Jiang, Y. Liu, Y. Chen,
Huang, X. Duan, Nano Lett. 2014, 14, 5590.
[256] G.-H. Lee, C.-H. Lee, A. M. van der Zande, M. Han, X. Cui, G. Arefe, C.
Nuckolls, T. F. Heinz, J. Hone, P. Kim, APL Mater. 2014, 2, 092511.
[257] N. Huo, J. Kang, Z. Wei, S.-S. Li, J. Li, S.-H. Wei, Adv. Funct. Mater. 2014,
24, 7025.
[258] Y. Xue, Y. Zhang, Y. Liu, H. Liu, J. Song, J. Sophia, J. Liu, Z. Xu, Q. Xu,
Z. Wang, J. Zheng, Y. Liu, S. Li, Q. Bao, ACS Nano 2016, 10, 573.
[259] N. Huo, J. Yang, L. Huang, Z. Wei, S.-S. Li, S.-H. Wei, J. Li, Small 2015,
11, 5430.
[260] N. Flöry, A. Jain, P. Bharadwaj, M. Parzefall, T. Taniguchi, K. Watanabe,
L. Novotny, Appl. Phys. Lett. 2015, 107, 123106.
[261] A. Pezeshki, S. H. H. Shokouh, T. Nazari, K. Oh, S. Im, Adv. Mater. 2016,
28, 3216.
[262] I. G. Lezama, A. Arora, A. Ubaldini, C. Barreteau, E. Giannini, M.
Potemski, A. F. Morpurgo, Nano Lett. 2015, 15, 2336.
[263] M. Long, E. Liu, P. Wang, A. Gao, H. Xia, W. Luo, B. Wang, J. Zeng, Y.
Fu, K. Xu, W. Zhou, Y. Lv, S. Yao, M. Lu, Y. Chen, Z. Ni, Y. You, X. Zhang, S.
Qin, Y. Shi, W. Hu, D. Xing, F. Miao, Nano Lett. 2016, 16, 2254.
[264] Y. Deng, Z. Luo, N. J. Conrad, H. Liu, Y. Gong, S. Najmaei, P. M. Ajayan,
J. Lou, X. Xu, P. D. Ye, ACS Nano 2014, 8, 8292.
[265] T. Hong, B. Chamlagain, T. Wang, H.-J. Chuang, Z. Zhou, Y. Xu,
Nanoscale 2015, 7, 18537.
[266] L. Ye, H. Li, Z. Chen, J. Xu, ACS Photonics 2016, 3, 692.
[267] S. Chuang, R. Kapadia, H. Fang, T. Chia Chang, W.-C. Yen, Y.-L. Chueh,
A. Javey, Appl. Phys. Lett. 2013, 102, 242101.
[268] J.-H. Ahn, M.-J. Lee, H. Heo, J. H. Sung, K. Kim, H. Hwang, M.-H. Jo,
Nano Lett. 2015, 15, 3703.
[269] S. Yang, C. Wang, C. Ataca, Y. Li, H. Chen, H. Cai, A. Suslu, J. C.
Grossman, C. Jiang, Q. Liu, S. Tongay, ACS Appl. Mater. Inter- faces 2016, 8,
2533.
[270] D. He, Y. Pan, H. Nan, S. Gu, Z. Yang, B. Wu, X. Luo, B. Xu, Y. Zhang,
Y. Li, Z. Ni, B. Wang, J. Zhu, Y. Chai, Y. Shi, X. Wang, Appl. Phys. Lett. 2015,
107, 183103.
[271] H.-C. Cheng, G. Wang, D. Li, Q. He, A. Yin, Y. Liu, H. Wu, M. Ding, Y.
Huang, X. Duan, Nano Lett. 2016, 16, 367.
[272] X. Hu, X. Zhang, L. Liang, J. Bao, S. Li, W. Yang, Y. Xie, Adv. Funct.
Mater. 2014, 24, 7373.
[273] Y. Guo, C. Liu, H. Tanaka, E. Nakamura, J. Phys. Chem. Lett. 2015, 6, 535.
[274] X. Duan, C. Wang, J. C. Shaw, R. Cheng, Y. Chen, H. Li, X. Wu, Y. Tang,
Q. Zhang, A. Pan, J. Jiang, R. Yu, Y. Huang, X. Duan, Nat. Nanotechnol. 2014, 9,
1024.
[275] C. Huang, S. Wu, A. M. Sanchez, J. J. P. Peters, R. Beanland, J. S. Ross, P.
Rivera, W. Yao, D. H. Cobden, X. Xu, Nat. Mater. 2014, 13, 1096.
[276] Y. Gong, J. Lin, X. Wang, G. Shi, S. Lei, Z. Lin, X. Zou, G. Ye, R. Vajtai,
B. I. Yakobson, H. Terrones, M. Terrones, B. K. Tay, J. Lou, S. T. Pantelides, Z.
Liu, W. Zhou, P. M. Ajayan, Nat. Mater. 2014, 13, 1135.
[277] M.-Y. Li, Y. Shi, C.-C. Cheng, L.-S. Lu, Y.-C. Lin, H.-L. Tang, M.-L. Tsai,
C.-W. Chu, K.-H. Wei, J.-H. He, W.-H. Chang, K. Suenaga, L.-J. Li, Science 2015,
349, 524.
[278] Y. Gong, S. Lei, G. Ye, B. Li, Y. He, K. Keyshar, X. Zhang, Q. Wang, J.
Lou, Z. Liu, R. Vajtai, W. Zhou, P. M. Ajayan, Nano Lett. 2015, 15, 6135.
[279] W. Feng, W. Zheng, X. Chen, G. Liu, W. Cao, P. Hu, Chem. Mater. 2015,
27, 983.
[280] H.-M. Li, D. Lee, D. Qu, X. Liu, J. Ryu, A. Seabaugh, W. J. Yoo, Nat.
Commun. 2015, 6, 6564.
[281] Y. Jin, D. H. Keum, S.-J. An, J. Kim, H. S. Lee, Y. H. Lee, Adv. Mater.
2015, 27, 5534.
[282] M. S. Choi, D. Qu, D. Lee, X. Liu, K. Watanabe, T. Taniguchi, W. J. Yoo,
ACS Nano 2014, 8, 9332.
[283] X. Yu, S. Zhang, H. Zeng, Q. J. Wang, Nano Energy 2016, 25, 34.
[284] M.-L. Tsai, S.-H. Su, J.-K. Chang, D.-S. Tsai, C.-H. Chen, C.-I. Wu, L.-J.
Li, L.-J. Chen, J.-H. He, ACS Nano 2014, 8, 8317.
[285] O. Lopez-Sanchez, E. Alarcon Llado, V. Koman, A. Fontcuberta i Morral,
A. Radenovic, A. Kis, ACS Nano 2014, 8, 3042.
[286] Y. Li, C.-Y. Xu, J.-Y. Wang, L. Zhen, Sci. Rep. 2014, 4, 7186.
[287] L. Hao, Y. Liu, W. Gao, Z. Han, Q. Xue, H. Zeng, Z. Wu, J. Zhu, W. Zhang,
J. Appl. Phys. 2015, 117, 114502.
[288] L. Wang, J. Jie, Z. Shao, Q. Zhang, X. Zhang, Y. Wang, Z. Sun, S.-T. Lee,
Adv. Funct. Mater. 2015, 25, 2910.
[289] Y. Zhang, Y. Yu, L. Mi, H. Wang, Z. Zhu, Q. Wu, Y. Zhang, Y. Jiang, Small
2016, 12, 1062.
[290] B. Li, G. Shi, S. Lei, Y. He, W. Gao, Y. Gong, G. Ye, W. Zhou, K. Keyshar,
J. Hao, P. Dong, L. Ge, J. Lou, J. Kono, R. Vajtai, P. M. Ajayan, Nano Lett. 2015,
15, 5919.
[291] X. Yuan, L. Tang, S. Liu, P. Wang, Z. Chen, C. Zhang, Y. Liu, W. Wang,
Y. Zou, C. Liu, N. Guo, J. Zou, P. Zhou, W. Hu, F. Xiu, Nano Lett. 2015, 15, 3571.
[292] S. Liu, X. Yuan, P. Wang, Z.-G. Chen, L. Tang, E. Zhang, C. Zhang, Y. Liu,
W. Wang, C. Liu, C. Chen, J. Zou, W. Hu, F. Xiu, ACS Nano 2015, 9, 8592.
[293] H. Zhang, X. Zhang, C. Liu, S.-T. Lee, J. Jie, ACS Nano 2016, 10, 5113.
[294] D. Li, R. Cheng, H. Zhou, C. Wang, A. Yin, Y. Chen, N. O. Weiss, Y.
Huang, X. Duan, Nat. Commun. 2015, 6, 7509.
[295] P. Gehring, R. Urcuyo, D. L. Duong, M. Burghard, K. Kern, Appl. Phys.
Lett. 2015, 106, 233110.
[296] Z. Xu, S. Lin, X. Li, S. Zhang, Z. Wu, W. Xu, Y. Lu, S. Xu, Nano Energy
2016, 23, 89.
[297] F. Xue, L. Chen, J. Chen, J. Liu, L. Wang, M. Chen, Y. Pang, X. Yang, G.
Gao, J. Zhai, Z. L. Wang, Adv. Mater. 2016, 28, 3391.
[298] S. Vélez, D. Ciudad, J. Island, M. Buscema, O. Txoperena, S. Parui, G. A.
Steele, F. Casanova, H. S. J. van der Zant, A. Castellanos- Gomez, L. E. Hueso,
Nanoscale 2015, 7, 15442.
[299] F. Liu, W. L. Chow, X. He, P. Hu, S. Zheng, X. Wang, J. Zhou, Q. Fu, W.
Fu, P. Yu, Q. Zeng, H. J. Fan, B. K. Tay, C. Kloc, Z. Liu, Adv. Funct. Mater. 2015,
25, 5865.
[300] M. Fontana, T. Deppe, A. K. Boyd, M. Rinzan, A. Y. Liu, M. Paranjape,
P. Barbara, Sci. Rep. 2013, 3, 1634.
[301] H. Jeong, H. M. Oh, S. Bang, H. J. Jeong, S.-J. An, G. H. Han, H. Kim, S.
J. Yun, K. K. Kim, J. C. Park, Y. H. Lee, G. Lerondel, M. S. Jeong, Nano Lett.
2016, 16, 1858.
[302] D. Jariwala, V. K. Sangwan, C.-C. Wu, P. L. Prabhumirashi, M. L. Geier,
T. J. Marks, L. J. Lauhon, M. C. Hersam, Proc. Natl. Acad. Sci. 2013, 110, 18076.
[303] K. H. Lee, T.-H. Kim, H.-J. Shin, S.-W. Kim, Adv. Mater. 2016, 28, 1793.
[304] J. Lin, H. Li, H. Zhang, W. Chen, Appl. Phys. Lett. 2013, 102, 203109.
[305] J. Miao, W. Hu, Y. Jing, W. Luo, L. Liao, A. Pan, S. Wu, J. Cheng, X. Chen,
W. Lu, Small 2015, 11, 2392.
[306] A. Sobhani, A. Lauchner, S. Najmaei, C. Ayala-Orozco, F. Wen, J. Lou, N.
J. Halas, Appl. Phys. Lett. 2014, 104, 031112.
[307] S. Najmaei, A. Mlayah, A. Arbouet, C. Girard, J. Léotin, J. Lou, ACS Nano
2014, 8, 12682.
[308] T. Hong, B. Chamlagain, S. Hu, S. M. Weiss, Z. Zhou, Y.-Q. Xu, ACS Nano
2015, 9, 5357.
[309] W. Wang, A. Klots, D. Prasai, Y. Yang, K. I. Bolotin, J. Valentine, Nano
Lett. 2015, 15, 7440.
[310] H. Li, Y. Anugrah, S. J. Koester, M. Li, Appl. Phys. Lett. 2012, 101.
[311] X. Gan, R.-J. Shiue, Y. Gao, I. Meric, T. F. Heinz, K. Shepard, J. Hone, S.
Assefa, D. Englund, Nat. Photonics 2013, 7, 883.
[312] X. Wang, Z. Cheng, K. Xu, H. K. Tsang, J.-B. Xu, Nat. Photonics 2013, 7,
888.
[313] N. Youngblood, C. Chen, S. J. Koester, M. Li, Nat. Photonics 2015, 9, 249.
[314] X. Liu, T. Galfsky, Z. Sun, F. Xia, E. Lin, Nat. Photonics 2015, 9, 30.
[315] S. Wu, S. Buckley, A. M. Jones, J. S. Ross, N. J. Ghimire, J. Yan, D. G.
Mandrus, W. Yao, F. Hatami, J. Vuckovic, A. Majumdar, X. Xu, 2D Mater.
2014, 1, 011001.
[316] S. Wu, S. Buckley, J. R. Schaibley, L. Feng, J. Yan, D. G. Mandrus, F.
Hatami, W. Yao, J. Vucˇkovic´, A. Majumdar, X. Xu, Nature 2015, 520, 69.
[317] S. Schwarz, S. Dufferwiel, P. M. Walker, F. Withers, A. A. P. Trichet, M.
Sich, F. Li, E. A. Chekhovich, D. N. Borisenko, N. N. Kolesnikov, K. S.
Novoselov, M. S. Skolnick, J. M. Smith, D. N. Krizhanovskii, A. I. Tartakovskii,
Nano Lett. 2014, 14, 7003.

You might also like