Next Article in Journal
Remote System Update for System on Programmable Chip Based on Controller Area Network
Next Article in Special Issue
Two-Dimensional Electronics and Optoelectronics: Present and Future
Previous Article in Journal
On the Generality of Codebook Approach for Sensor-Based Human Activity Recognition
Previous Article in Special Issue
High Throughput Characterization of Epitaxially Grown Single-Layer MoS2
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Recent Advances in Electronic and Optoelectronic Devices Based on Two-Dimensional Transition Metal Dichalcogenides

Department of Physics, Michigan Technological University, 1400 Townsend Drive, Houghton, MI 49931, USA
*
Author to whom correspondence should be addressed.
Electronics 2017, 6(2), 43; https://doi.org/10.3390/electronics6020043
Submission received: 16 March 2017 / Revised: 1 May 2017 / Accepted: 12 May 2017 / Published: 2 June 2017
(This article belongs to the Special Issue Two-Dimensional Electronics and Optoelectronics)

Abstract

:
Two-dimensional transition metal dichalcogenides (2D TMDCs) offer several attractive features for use in next-generation electronic and optoelectronic devices. Device applications of TMDCs have gained much research interest, and significant advancement has been recorded. In this review, the overall research advancement in electronic and optoelectronic devices based on TMDCs are summarized and discussed. In particular, we focus on evaluating field effect transistors (FETs), photovoltaic cells, light-emitting diodes (LEDs), photodetectors, lasers, and integrated circuits (ICs) using TMDCs.

1. Introduction

The miniaturization of silicon-based transistors has led to unprecedented progress in smaller and faster electronic devices including smartphones and tablet computers. Further miniaturization of transistors is hindered by some fundamental limits including high contact resistance, short channel effects, and high leakage currents [1,2,3]. Tremendous efforts have been attempted to overcoming these limitations using novel device architecture [4,5]. In addition, silicon-based devices are not applicable for flexible electronics. All the shortcomings of silicon-based transistors have led to the motivation for the search of new device concepts and alternative materials. For example, “transistors without using semiconductors” have been proposed using boron nitride nanotubes functionalized with gold quantum-dots (QDs-BNNTs) [6,7] and graphene–BNNT heterojunctions [8]. All these materials are applicable for flexible electronics. On the other hand, two-dimensional transition metal dichalcogenides (2D TMDCs) offer several attractive features for next-generation transistors, including high electron mobility (up to ~1000 cm2 V−1·s−1 for MoS2), atomically thin and flexible. In addition, TMDCs can be free of dangling bonds on their surfaces, and offer many interesting optical properties for their potential application in optoelectronic devices.
Although research publication on TMDCs was started around the 1960s, publication activity has increased significantly since 2014. According to our analysis in Web of ScienceTM, a total of 3829 journal articles were published up to 2016, with 33.8%, 21.5%, 12.5%, and 5.5% of these articles were published in 2016, 2015, 2014, and 2013, respectively. Based on these recent publications, we will discuss the application of 2D TMDCs in field effect transistors (FETs), optoelectronic devices (photovoltaic cells, light-emitting diodes, and lasers), and integrated circuits (ICs). Interested readers may refer to a recent review on the same topics [9]. For the convenience of the readers, we summarize a series of TMDCs along with the nature of their band gaps in Table 1.

2. Field Effect Transistors

Field-effect transistors (FETs) are among the basic devices for modern electronic circuits. A typical FET modulates the conductivity of the semiconductor channel between the source (S) and the drain (D) terminals by controlling the applied voltage on the gate (G) electrode. Some of the important figures of merit for FETs are defined here. First, a good FET should have a low subthreshold swing/slope:
S S = ( d ( log I D S ) d V G ) 1
which is defined as the gate voltage (VG) required to change the source–drain current (IDS) by one order of magnitude (in the unit of volt per decade). FETs are treated as switches that offer a high conductivity of the channel (ON state), and vice versa (OFF state). Hence, the on/off ratio is another crucial factor in measuring the performance of FETs. In addition, mobility is important for high-performance FETs. In 2D TMDCs, the transport of charge carriers is confined to the plane of the materials. The mobility of carriers is related to the scattering by
μ = e τ m
where τ is the scattering time, and m* is the effective mass of in-plane electron. There are following scattering mechanisms: (1) electron–phonon scattering, including longitudinal acoustic (LA), transverse acoustic (TA) [31,32], polar (longitudinal) optical (LO), and homopolar optical phonons [33]; (2) electron–electron (Coulomb) scattering [34] and charged impurities scattering [35]; (3) surface (interface) roughness Coulomb scattering [34] and phonon scattering [36]; and (4) short-range scattering, which is from the defect and dislocation of the lattice [35,37,38]. The degree to which these mechanics affect the mobility can be evaluated by Matthiessen’s rule:
1 μ = 1 μ 1 + 1 μ 2 + 1 μ 3 + 1 μ 4 .
Finally, Flicker noise is a factor that limits the performance of the FETs. This is a type of electronic noise with a frequency spectrum such that the power density spectrum is inversely proportional to frequency, f, and is therefore referred as 1/f noise. Flicker noise occurs in almost all electronic devices and often shows up as the resistance fluctuation. It is generally related to the impurities in the conductive channel and generation/recombination noise in transistors due to memory of the base current [39].

2.1. Transistors with Multilayered TMDCs

Although the electrical properties and electronic structures of TMDCs were studied in the 1960s [33], their device applications have been limited until 2004. One of the earliest works in which TMDCs were used for FETs was reported by comparing WSe2 with single-crystal Si metal oxide FETs. A p-type conductivity as high as 500 cm2·V−1·s−1 was demonstrated at room temperature with a 104 on/off ratio at 60 K [40]. Thus, a further study based on multilayer MoS2 FETs with back-gate configuration was conducted, but a rather low mobility (0.1–40 cm2·V−1·s−1) was demonstrated [41,42,43]. Apparently, the reported mobility from multilayered MoS2 was quite lower as compared to those reported for bulk MoS2, where mobility is often ranged from 100 to 260 cm2·V−1·s−1 [33]. Table 2 summarizes the performances of FETs constructed by various TMDCs. It is noted that the mobility of graphene was much higher, up to 10,000 cm2·V−1·s−1, even when using SiO2 as the gate materials [44].
Later, by using high-k dielectrics (50-nm-thick Al2O3), FETs with a back-gate configuration were reported with relatively high mobility (>100 cm2·V−1·s−1) and on/off ratio (>106) based on multilayer MoS2 (Figure 1a). The width and length of the device were 4 μm and 7 μm, respectively. Thirty-nanometer-thick multilayer MoS2 was deposited on a silicon substrate that was coated with a 50-nm-thick Al2O3 layer. As shown in Figure 1b,c, the near-ideal subthreshold slope/swing of 80 mV/dec and robust current saturation over a large voltage window are demonstrated [46]. Accumulation mode occurred at a positive gate bias, and the window of depletion mode occurred at negative gate bias. At large negative gate biases, the drain current recovers and forms an inversion mode. Furthermore, FETs with Schottky contacts and Al2O3 gate oxide (15-nm-thick) on the top of MoS2 channel with back-gated configuration was shown to have mobility as high as 700 cm2·V−1·s−1 at room temperature [49]. Liu et al. have further demonstrated that the field effect mobility of multilayer (20 layers) MoS2 FETs exceeds 500 cm2·V−1·s−1 due to the smaller bandgap (thus a smaller Schottky barrier) [48], compared with monolayer MoS2 FET within the same top-gated configuration [82].
The dependence of carrier mobility on temperature in multilayer MoS2-based FETs is due to different scattering mechanisms as shown in Figure 2. At low temperature, the carrier mobility is limited by charged impurity scattering. The mobility is determined by the combined effect of the homopolar phonon and the polar-optical phonon scatterings at room temperature [46]. Although the high-κ dielectric materials (e.g., Al2O3 and HfO2) may screen Coulomb scattering from charged impurities [49], the complete recovery of the intrinsic phonon-limited mobility has not been observed in high-κ dielectrics encapsulated TMDCs devices at room temperature.
Another limiting factor of mobility is the substrates. It was shown that trapped charges in the interface of MoS2–SiO2 would lead to carrier localization in the bilayer and trilayer MoS2 channel [83]. Considerable mobility improvement was reported in multilayer MoS2 on PMMA dielectric in comparison with MoS2 on SiO2, which was attributed to the reduced short-range disorder or long-range disorder at the MoS2/PMMA interface [52]. Other phenomena, such as thickness-independent SS and hysteresis as the gate swept, have been attributed to the interface traps between the bottom layer of WS2/MoTe2 and SiO2 substrate [64,68]. Besides, the presence of a low-energy optical phonon mode in SiO2 (∼60 meV) may also cause non-negligible surface polar optical scattering. A two to three times mobility improvement is consistently observed in MoSe2 FETs on parylene-C substrates compared to SiO2 [55]. WSe2 FETs with parylene top-gate dielectric have demonstrated high room-temperature mobility up to 500 cm2·V−1·s−1 [40].
Low carrier mobility in TMDC devices might be due to the shortcomings of two-terminal measurements. A mobility of 306.5 cm2·V−1·s−1 was extracted from the same multilayer MoS2 device when using a four-terminal configuration, which is considerably greater than 91 cm2·V−1·s−1 measured by a two-terminal configuration [51]. This results indicated that the intrinsic carrier mobility on the SiO2 substrate is significantly greater than the values previously reported in MoS2 FETs [42]. The analogous configuration also applied to FETs based on MoS2 [52], MoSe2 [55], MoTe2 [59], WS2 [68], and WSe2 [40]. Therefore, future studies of TMDCs devices should use the four-terminal measurement.
Table 2 summarizes the performances of FETs constructed by various multilayered TMDCs. By comparing the carrier mobility, the on/off ratio, and the subthreshold swing of these devices, we found that devices constructed by CVD MoS2 [45], ReS2 [79], and ReSe2 [81] exhibited very low mobility, which are two or three orders of magnitudes less than those based on exfoliated TMDCs. The relatively low mobility reported in these earlier literatures is likely due to structural defects, such as grain boundary [84]; these grain boundaries act as short-range scattering centers to suppress the carrier transport [85]. After the CVD synthesis conditions were optimized to reduce the grain boundary, the mobility increased to a level similar to that based on exfoliated TMDCs [50,55,70,71]. It was reported that the threshold voltage for devices with exfoliated MoS2 was higher than that based on CVD MoS2 [86]. The possible reasons could be the density of unintentional doping and the Scotch tape residue-induced charges in the exfoliated samples.
Furthermore, as compared to the low-resistance ohmic contacts, the Schottky contacts between the TMDCs and gates are one of the reasons for low carrier mobility. It is believed that the nature of contact is more important than the properties of the TMDCs [87]. By using a non-contact method based on THz-probe spectroscopy, an intrinsic mobility of MoS2 was found to be approaching 4200 cm2·V−1·s−1 at 30 K, which is an order of magnitude greater than any other previously reported values in multilayer TMDCs devices [88].
The intrinsic current distribution in FETs based on multilayer MoS2 has been investigated. Here, the 13-layer MoS2 FETs (8 nm in total thickness) were treated as consecutive resistors in a resistor network model that is based on the Thomas−Fermi charge screening and interlayer coupling effects. The authors map the current distribution among the individual layers of the multilayer 2D systems. Result suggests the existence of a centroid of current distribution or the so-called “HOT-SPOT” in multilayer MoS2 FETs [89]. Figure 3 shows the current distribution in the multilayer MoS2 FETs at different gate bias conditions. As shown, the current mainly travels through the upper layer that is closer to the source and the drain electrodes. The “HOT-SPOT” still moved toward the top layers even when the gate bias was increased. Hence, the large effective interlayer resistance and charge screening effect will limit the carrier concentration of multilayer FETs.
Recently, FETs based on multilayer MoS2 with an intrinsic gain over 30, an intrinsic cut-off frequency up to 42 GHz (Ft), and a maximum oscillation frequency up to 50 GHz (Fmax) were fabricated based on optimum contacts and device geometry [90]. These performances are better than those reported so far for MoS2 transistors [91]. Multilayer MoS2 FETs with a 1 nm physical gate length using a single-walled carbon nanotube (SWCNT) as the gate electrode were demonstrated. These ultrashort devices exhibit excellent switching characteristics with a near-ideal subthreshold swing of 65 mV/dec and an on/off current ratio of 106. Simulations show an effective channel length of ~3.9 nm in the off state and ~1 nm in the on state [53].
Finally, while MoS2 FETs are one of the most studied TMDCs, MoSe2 FETs have been reported to have a low noise characteristic with a Hooge parameter of 3.75 × 10−3 [92], which is similar or lower than MoS2 FETs [93,94]. On the other hand, WS2 was once predicted to offer the best transistor performance (the highest on-state current density and mobility) among all TMDCs [95]. However, only a few WS2 FETs have been reported to have a performance comparable to that of MoS2 [62]. In addition, MoTe2 offers a favorable bandgap (Table 1) in the near-infrared range. Therefore, MoTe2 FETs are an appropriate candidate to substitute Si and has great potential for high-performance optoelectronic devices [59,63,64]. Finally, ReS2 and ReSe2 are TMDCs with a direct bandgap in monolayer, multilayer, and bulk forms, as summarized in Table 1 and Table 2. They have recently attracted much interest and have been applied to optoelectronic devices.

2.2. Transistors with Monolayer TMDCs

Table 3 summarizes performances of FETs based on monolayer of TMDCs in various gate configurations. As shown, the electron mobility is highest for devices based on exfoliated MoS2. Apparently, other TMDCs offer much lower mobility. In addition, the mobility of CVD MoS2 is also significantly lower.
Theoretical simulation was applied to predict the performance limits of transistors based on monolayer MoS2. It is known that optical phonon scattering is the intrinsic scattering mechanism that dominates at room temperature. On the other hand, acoustic scattering is dominated at lower temperature (T < 100 K) [118,119]. Figure 4 shows the temperature dependence of electron mobility in freestanding monolayer MoS2 as predicted by first principle calculations. As reported, the mobility (μ) is expected to follow the T−γ temperature (T) dependence with γ = 1.69. It was predicted that the mobility of monolayer MoS2 can reach 410 cm2·V−1·s−1 at room temperature [120]. This theoretical limit has become the reference for experimental efforts. Indeed, among the TMDCs, the mobilities of MoTe2 (2526 cm2·V−1·s−1), WS2 (1103 cm2·V−1·s−1), and WSe2 (705 cm2·V−1·s−1) were predicted to be greater than the room-temperature phonon-limited electron mobility in the monolayer [121], as a result of low effective mass. However, as shown in Table 3, the reported experimental mobility to date is much lower than that found via theoretical prediction.
The pioneer work on FETs based on monolayer MoS2 was demonstrated by Radisavljevic et al. in 2011, as shown in Figure 5 [82]. The reported performance includes an excellent on/off current ratio (~108), greater mobility (>200 cm2·V−1·s−1), an ideal low subthreshold swing (74 mV/dec), and smaller off-state currents (25 fA/μm). Here, a mechanical exfoliated monolayer MoS2 (0.65 nm) was used as the transport channel and was covered by 30-nm-thick HfO2, which served as the top-gated dielectric layer. The same configuration was also used in FETs with channels made of monolayer WSe2 covered by ZrO2 dielectrics. These FETs also exhibited high room-temperature mobility (~250 cm2·V−1·s−1), a good subthreshold swing of ~60 mV/dec, and a high on/off ratio of 106 [115].
It is believed that the encapsulation of 2D-TMDCs by using high dielectric materials (e.g., Al2O3 and HfO2) is critical for achieving high mobility. Authors have suggested that the introduction of high-k dielectrics would strongly dampen the Coulombic scattering in 2D materials due to the dielectric screen [122]. This mobility increase in conjunction with dielectric deposition was also observed in multilayer MoS2 [46] and monolayer MoS2 with polymer electrolytes [100]. The scattering rate is related to the scattering mean free path and the impurity density (i.e., the impurity spacing). Therefore, the charge impurity must be appropriate such that the scattering mean free path is on the same or small order of the phonon mean free path. Therefore, a minimum impurity concentration of 5 × 1011 cm−2 is need to dominate phonon scattering [120], which corresponded to the heavy doping by using top-gated dielectric materials. The similar mobility improvement due to such dielectric engineering has been reported earlier for graphene [123,124,125].
On the other hand, the effects of gate configuration on charge mobility were also studied for FETs based on monolayer MoS2. Top-gated geometry allows for a lower turn-on voltage, and the integration of multiple devices on the same substrate. Top gate configuration also had the effect of quenching the homopolar phonon mode, which has contributed to a mobility enhancement of 70 cm2·V·s−1 at room temperature [120]. However, the measured exponent γ factors in the temperature dependence mobility, μ = T−γ, are between 0.3 and 0.78 for top-gated devices [126]. These values are much smaller than the theoretically predicted value of 1.52 for monolayer MoS2 [120] or 2.6 for TMDC bulk crystals [33]. These results suggest that other mechanisms, rather than the quenching of the homopolar phonon, such as the surface roughness scattering created during synthesis, exfoliation, and transferring processes, are affecting the mobility of monolayer MoS2 in the top-gated devices [127].
In addition, the Schottky barrier height between the electrodes and 2D TMDCs is known to affect carrier mobility of devices. The metals used for electrodes are claimed to define the height of the Schottky barrier associated with their work function. However, Theoretical calculations have shown 2D-TMDCs tend to form high Schottky barriers (0.03 eV to 0.8 eV) with common metals [128,129,130,131]. Several experimental practical attempts by inserting 2D layers (Graphene [132,133], hBN [134], Nb doped TMDCs [135], MoO3 [136]) or thin film (MgO [137], TiO2 [138]) between TMDCs and metals have been made to reduce Schottky contacts. Besides the handling of TMDC/metal junctions, chemical/electostatic-doping [69,115] and phase-patterning [139,140,141,142] are two successful methods of modifying the electrical properties of channel by reducing contact resistance. It was also reported that in situ annealing at 120 °C for 20 h in vacuum (~10−6 mbar) could increase the mobility due to the reduction of Schottky barrier. By this approach, an intrinsic mobility of 1000 cm2·V−1·s−1 were observed for both monolayer and bilayer devices based on MoS2 [104].
Theoretical simulation has been investigated for the performance limits of an FET with a 15-nm-long MoS2 gate length [96]. It has been reported that FETs based on MoS2 can offer a subthreshold swing as low as 60 mV/dec and a current on/off ratio as high as 1010. The ballistic regime of this device can be as high as 1.6 mA/μm, and the drain-induced barrier lowering (DIBL) is as small as 10 mV/V even with a very short channel length [96]. Thus, TMDCs transistors have the potential to overcome the short-channel effects because of the enhanced gate effect due to their atomic-scale thickness. A similar simulation was reported based on a two-gate MOSFET model with 2D-Silicon and different 2D-TMDCs. Among them, only the monolayer WS2 transistor outperform the 2D silicon and other monolayer TMDC devices in terms of ON-current by about 28.3% [95]. However, these predictions did not consider the effects of edges on the transport properties and therefore require experimental verification.
It should be noted that high transconductance FETs based on monolayer MoS2 with full-channel gates were demonstrated experimentally [103]. As shown in Figure 6a, at low bias Vds = 1 V, the drain–source conductance is close to 0 (gds < 2 μS/μm, g d s = d I d s d V d s ), which means the channel current saturation occur. As shown in Figure 6b, these FETs offer high transconductance g m = d I d s d V t g (maximum transconductance gm = 34 μS/μm for Vds = 4 V). Furthermore, it was demonstrated that the electrical breakdown of the monolayer MoS2 was recorded at a current density of 5 × 107 A/cm2, exceeding the current-carrying capacity of Cu by 50 times.
The performance of monolayer TMDCs are also sensitive to random perturbations in the local environment. Current hysteresis under different ambient condition and light illumination was reported for FETs based on monolayer MoS2. It was reported that current hysteresis increase steadily with the increase in humidity (RH), as shown in Figure 7a. It was proposed that such a hysteresis is due to the trapping states induced by the adsorption of water molecules on the surface of MoS2. Similar hysteresis was observed when the device was under white light illumination, as shown in Figure 7b. This is attributed to the improvement of carrier concentration in the conduction channel due to photosensitivity [99]. This work indicates that controllable hysteretic behavior in MoS2 FETs has the potential for humidity sensors and non-volatile memory devices.
The sensitivity of 2D TMDCs to the local environment has also contributed to flicker noise. The 1/f noise from un-encapsulated FETs based on monolayer MoS2 was explained by the Hooge parameter ranging between 0.005 and 2 in vacuum (<10−5 Torr) [143]. The noise amplitude was reported to reduce by an order of magnitude after annealing, revealing the significant influence of atmospheric adsorbates on the charge transport [144]. On the other hand, high-k dielectric for top gate configuration is known to reduce the drain current noise in both monolayer and multilayer MoS2 FETs [94,145]. Additionally, the observation of low frequency generation-recombination noise at low temperature could be due to charge traps in the underlying SiO2 substrate or mid-gap states in the monolayer MoS2 [143,146]. On the other hand, FETs based on layered 2H-MoTe2 flakes were investigated for their low frequency noise in both vacuum and air [147]. Similar experiments were also performed using as-fabricated and aged MoS2 transistors [148]. The measurements demonstrate that the flicker noise is determined by an intrinsic TMDCs conducting channel rather than the contact barriers. As a result, these noise metrics are expected to provide a unique diagnostic tool for researchers as they develop high-performance sensing applications.
The effects of substrates on the performance of FETs constructed by monolayer CVD TMDCs were investigated [70,106,108]. Results suggest that these FETs offer similar electrical performance to devices based on exfoliated TMDCs. This may imply that FETs based on CVD TMDCs can be fabricated on flexible substrates, enable their applications in flexible, transparent, 2D electronic devices. The development of CVD synthesis methods for obtaining large areas of TMDCs has become important for wafer-scale device fabrication.
Finally, FETs based on monolayer TMDC alloys were also investigated. These FETs were based on an MoWSe2 monolayer (Mo1-xWxSe2, 0 ≤ x ≤ 1), mechanically exfoliated by Scotch tape. It was demonstrated that these FETs exhibit n-type transport behavior with on/off ratios >105 [149].

2.3. Ambipolar Transistors with TMDCs

Ambipolar transport was demonstrated in electric double-layer transistors (EDLTs) based on thin-flake MoS2 (10 nm) using an ionic liquid as the gate to reach extremely high carrier densities of 1 × 1014 cm−2 [150]. The charge transfer curves of such ambipolar devices are shown in Figure 8b. For EDLTs using bulk and thin MoS2, the IDS was increased with positive VG, corresponding to the behavior of an n-type semiconductor transistor. In contrast to its commonly known property as an n-type semiconductor in Table 2, when the VG turn to negative bias, an obvious p-type transport was observed to start at −1 V in the EDLT made of thin flake MoS2. There was no significant change in current in the EDLT made of bulk MoS2. This phenomenon indicates that hole transport is comparable to electron transport in thin-flake MoS2. Hall effect measurements also revealed the mobility of 44 and 86 cm−1/V−1s−1 for electron and hole, respectively. The hole mobility is even twice the value of the electron mobility. However, the on/off ratio in this device was just >100, much lower than those reported for FETs based on monolayer MoS2 [82], mainly because of the finite off-current passing through the thin flakes. The greater hysteresis at lower temperature (220 K) in Figure 8a was attributed to the slow motion of ions. A stable p–n junction was also found in the MoS2 EDLTs [151].
The electrical double layers (EDLs) consist of a narrow spatial charge doublet that simulates a capacitor to accumulate carriers. The use of EDLTs help to investigate the ambipolar transport in TMDCs, such as WS2 [20,66], MoSe2 [152], WSe2 [69,116,117], and MoTe2 [62]. In addition, the EDLT channel was shown to form at the interface of TMDC/ionic liquid, where the trapping state between the TMDCs and the substrates was reduced. As a result, the similar ambipolar behavior were also reported in conventional FETs based on TMDCs on dielectric materials (e.g., MoS2 on PMMA [52], WS2 on Al2O3 [65], and WSe2 on hexagonal boron nitride (h-BN) [153]).
Ambipolar EDLTs based on multilayer WSe2 with a high on/off ratio >107 at 170 K and large carrier mobility (electron mobility to 330 cm2·V−1·s−1 and hole mobility to 270 cm2·V−1·s−1) at 77 K were reported [69]. As shown in Figure 9, it was obtained by using a low-resistance ionic liquid gate with hBN encapsulation and graphene contacts on the WSe2. It was suggested that a drastic reduction of the Schottky barriers between the channel and the graphene contact electrodes by ionic liquid gating is needed in order to observe the intrinsic, phonon-limited conduction in both the electron and hole channels. However, the EDLTs still require much effort in atomic 2D-TMDC devices due to the integration, reliability, and low operating speed.
In the case of conventional FETs, another method is using electrode metals with different work functions to control the transistor polarity. If the Fermi level of the electrode is close to the conduction band of the semiconductor, the electron injection becomes predominant. While the Fermi level is close to the valence band of the semiconductor, it translates into a hole injection. For example, due to the Fermi level of most metals close to the conduction band of MoS2, the carrier of MoS2 FETs were dominated by electrons [49]. With the exception of Pd [154] and Pt [131], which acted as both source and drain electrodes, MoS2 transistors show p-type behavior. Furthermore, WSe2 was found to exhibit ambipolar behavior by using different metals as the source and drain electrodes [155]. Alternatively, chemical doping on the surface of TMDCs (e.g., NO2 on WSe2 [115], K on WSe2 [156]) by thinning the Schottky barrier width for carrier injection was also performed to study ambipolar behavior. However, the polarity of FETs was fixed after device fabrication.
The ambipolar transport were also discovered in MoSe2 [56], MoTe2 [58], and WSe2 [153,157,158,159] under different electrostatic fields. This phenomenon was attributed to the formation of two back-to-back Schottky barriers rather than only the channel conductance. Thus, transistors can function as ambipolar transistors, and the transistor polarity can be reversed by an electric signal. A doping-free ambipolar transistor made of multilayer MoTe2 is proposed in which the transistor polarity (p-type and n-type) is electrostatically controlled by dual top gates as shown in Figure 10. One of the gates was used to determine the transistor polarity, while the other gate was used to modulate the drain current [160].
Overall, the demonstration of both n-type and p-type transport will be useful for applications that are more complicated such as CMOS logic circuits and p–n junction optoelectronic devices.

2.4. Transistors with Vertical Hetero-Structures

Due to the relatively weak van der Waals bonding between layers of TMDCs, and the absence of surface dangling bonds, the stacking of TMDCs layers and other 2D materials multilayers can energetically form heterostructures or heterojunctions [161,162]. Actual devices have been demonstrated as summarized in Table 4. For example, vertical heterojunction devices consisting of graphene (Gr) and 2D-TMDCs with appropriate bandgaps were suggested to resolve the low on/off-current ratio of graphene FETs and to achieve a high on–current density at extremely low operating voltage, which is difficult to accomplish in TMDC FETs. The use of graphene in vertically stacked devices was demonstrated as allowing electric field to penetrate through the heterostructures [163]. These vertical heterojunctions created Schottky barriers at the interfaces, where the TMDCs functioned as semiconductor instead of insulator (e.g., hBN) tunneling [164], which achieved a greater channel current while retaining a high on/off ratio [165,166,167,168].
As summarized in Table 4, vertical Gr/TMDCs heterostructures have been fabricated since 2012 [165,166,167,168,169,170,171,172,173,174,175,176,177,178,179,180,181]. For example, vertical Gr/MoS2 FETs can achieve a high on/off ratio above 105 with on–current density as high as 5 × 104 A/cm2 [171]. In addition, a similar vertical FET based on Gr/WS2/Gr structure was reported with an on/off ratio as high as 106 even at room temperature [166]. At lower temperature (180 K), an extremely high on/off ratio of 5 × 107 was obtained in Gr/WSe2 FETs [175].
Unlike n-type semiconductors, MoS2, MoTe2, and WSe2 show p-type behavior, as summarized in Table 2 and Table 3. Hence, multiple MoS2/WSe2 vertical heterostructures have been reported as p–n junctions [182,183,184]. A heterojunction based on p-type black phosphorus (BP) and n-type monolayer MoS2 has also been demonstrated [185].
Although the study of transistors with vertical heterojunctions is still in its infancy, unique electrical and optical junction properties are expected to be found by combining various 2D materials with different work functions, bandgaps, and electron affinities. In fact, in addition to FETs, various TMDCs and other 2D materials heterojunction-based devices have recently been reported, including solar cells, photodetectors, LEDs, and memory devices, which will be discussed in following chapters.

3. Optoelectronic Devices

Optoelectronic devices are important to generate, detect, and control light, including photovoltaic devices (solar cells), light-emitting diodes (LEDs), photodetectors, and lasers. The electronic band structures of semiconductors decide for their possibility to absorb and emit light. For indirect bandgap semiconductors, additional phonons must be absorbed or emitted to conserve the momentum, which is much less efficient. Almost all MoX2 and WX2, TMDCs (X is a chalcogen atom) are expected to follow a similar indirect to direct bandgap transformation with decreasing layer thickness. The bandgap of these TMDCs covering an energy range of 1–2.1 eV, as shown in Table 1, are therefore especially suitable for optoelectronic applications [186]. Because of their atomically thin and strong stiffness [187], TMDCs are promising for flexible and transparent optoelectronics. As summarized in Figure 11, TMDCs offer good mobility in between those of organic materials and III–V semiconductors. A band gap near the Shockley–Quessier limit (∼1.3 eV) and high mobilities are expected to lead to high-efficiency photovoltaic devices [188].
When irradiated with photons with an energy greater than that of the band gap, free electron/hole or bound electron–hole pairs (excitons) will be generated in the depletion region of semiconductors. The applied voltage or built-in electrical field could separate the bound excitons to generate flows of charge carriers that are called photocurrents. Hence, the p–n junction is the functional element of many optoelectronic devices, including solar cells, LEDs, and photodetectors. Devices with a combination of p-type Si and n-type MoS2 were studied [189,190]. A p–n junction diode has been realized in MoS2 EDLTs [151]. Similarly, the p–n junctions based on monolayer WSe2 by electrostatic tuning were also reported [153,157,158,159].

3.1. Solar Cells

In Figure 12, TMDCs (MoS2, MoSe2, and WS2) can absorb up to 5–10% incident sunlight in monolayer (thickness less than 1 nm). The sunlight absorption is one order of magnitude greater than general semiconductors such as GaAs and Si [176]. Early to 1997, Gourmelon et al. first reported the use of MoS2 and WS2—nanostructure absorbers in solar cells [191]. Photosensitization of TiO2 were sensitized with WS2 nanosheets (5 nm thick), which act as a stable absorber material [192]. Similarly, a bulk heterojunction (BHJ) solar cell made of TiO2 nanoparticles and composites of monolayer/multilayer MoS2 nanosheets with poly(3hexylthiophene) (P3HT) was demonstrated with 1.3% photo conversion efficiency [193]. The asymmetric Schottky junctions were demonstrated in a metal/MoS2/metal structure, which resulted in photovoltaic devices with 0.7∼2.5% power conversion efficiency (PCE) [154,194]. A greater work function difference between metals and n-type MoS2 would generate a greater electric field in the depletion region of MoS2. Recently, the p–n junction diodes based on TMDCs themselves were fabricated and achieved a light PCE of ~0.5%. When the device was operated as a photodiode (Figure 13), a photocurrent of 29 pA was obtained, which translates into a photo-responsivity of 16 mA/W [157]. Furthermore, the monolayer MoS2 on p-type Si substrates formed a p–n junction and operated in tandem mode with an external quantum efficiency of 4.4% [189].
Highly efficient photocurrent generation was also demonstrated in vertical heterostructures. Studies on Schottky junction solar cells consisting of a graphene–MoS2 stack suggested a power conversion efficiency of 0.1–1% [176], while p–n junctions of MoS2/WS2 [176] and MoS2/WSe2 [182] were 0.4–1.5% and 0.2%, respectively. In addition, a special vertical p–njunction made of p-type MoS2 (CHF3 plasma treated) and n-type MoS2 was fabricated successfully. This MoS2 heterostructure-based solar cell achieved up to 2.8% PCE under AM1.5G illumination [195].
However, the efficiency of solar cells based on ultrathin TMDCs is limited by the loss of absorption under the thickness limitation [196]. Calculations suggest that a monolayer of TMDCs could absorb as much sunlight as 50 nm of Si and generate electrical currents as high as 4.5 mA/cm2 [176]. To improve the absorption of light, plasmonic enhancement technics were introduced by decorating Au nanoparticles on MoS2 nanosheets [197]. The Au nanoparticles can produce an enhanced local optical field on an MoS2 phototransistor device. A similar phenomenon was also investigated in the WS2/graphene and WSe2/graphene heterostructures, leading to enhanced photon absorption [177,178]. Alternatively, multi-junction solar cells with different bandgaps can convert different portions of the solar spectrum to reduce absorption loss [196,198]. On the other hand, a continuous bandgap tuning by strain was demonstrated [199]. A photovoltaic device made from a strain engineered MoS2 monolayer will capture a broad range of the solar spectrum and concentrate charge carriers. These structures can also potentially be constructed using different TMDCs with varying bandgaps, which range from the visible to the near-infrared (NIR).

3.2. Light-Emitting Diodes

Light-emitting diodes (LEDs) is another type of application based on p–n junctions. Electrons and holes are separated by doping material into a p–n junction. They recombine to release energy as photons in response to electrical bias voltages. This effect is called electroluminescence (EL). Table 5 summarizes performances of LEDs constructed by various TMDCs. As direct bandgap materials (1–2 eV), monolayer 2D-TMDCs are ideal for ultrathin and flexible light-emitting layers. One of the earliest light emission studies on TMDCs was one in which MoS2/Au nano-contacts were stimulated by scanning tunneling microscopy (STM) [200]. EL behavior was also found from exfoliated SnS2 incorporated into a composite polymer semiconductor, but MoS2 resulted in no light emission [201].
The result of EL in MoS2 was reconfirmed by a monolayer MoS2 FET, shown in Figure 14. A Schottky barrier between MoS2 and Cr/Au was formed with a height of 100 to 400 meV. By comparing the absorption, photoluminescence (PL) and EL of the same MoS2 monolayer, they all involved the same energy state at 1.8 eV (685 nm) [202]. WS2 were also used to fabricate light-emitting transistors based on similar Schottky barrier between TMDCs and metals [20].
More LEDs based on p–n junctions have been shown to have a high light emission efficiency (~1%). Homojunction-based LEDs formed by electrostatic doping of monolayer WSe2 resulted in a maximum EL efficiency of 1% [157,158,159]. However, the difficulty in doping has prevented the fabrication of LEDs by monolayer TMDCs. Moreover, based on the type II heterojunction design as previous mentioned in Section 3.1, excitonic EL was observed via a n-type monolayer MoS2 and p-type Si heterojunction [189]. In addition, an EL spectrum in a vertically stacked MoS2/WSe2 heterojunction (n-type for MoS2 and p-type for WSe2) was also obtained with a 12% external quantum efficiency [184].
Furthermore, a new design of LEDs was introduced by stacking graphene (electrode), hexagonal boron nitride (hBN) (tunneling junction), and various semiconducting monolayers (light emitter) into complex sequences. These devices have exhibited an extrinsic quantum efficiency of nearly 10%, and the emission can be tuned over a wide range of frequencies by appropriately choosing and combining 2D TMDCs [204,205]. However, as compared to the commercial organic LEDs with an emission efficiency (external quantum efficiency) of 15–40% [206] and of ~50% in LED light bulbs, there are significant needs to further improve the performance TMDC-based LEDs. For example, more investigation on doping and surface engineering, encapsulation, and device design for band structure engineering are desired to further improve the efficiency of these LEDs [188].
Recently, LEDs have combined vertical tunnel junctions, and lateral p–n junctions have also been exploited. Such lateral LED devices were constructed by two graphene transparent electrodes, two hBN tunnel barriers, and one monolayer WSe2. Two separated back-gates were utilized to electrostatically define the p–i–n junction in WSe2 [203,207]. A single photon emitter in a WSe2 LED based on a p–i–n junction was reported; a narrow EL peak (width < 300 μeV) at 1.704 eV (728 nm) under low current (~100 nA) and low temperature (~5 K) was observed [207].

3.3. Photodetectors

Photodetectors are light sensors with p–n junctions that convert photons into electrical current. The junction is usually covered by a window with anti-reflective coatings. The absorption of photons will create electron–hole pairs in the depletion region. Photodiodes and phototransistors are two major examples of such photodetectors. In terms of frequency response, photodiodes are much faster than phototransistors (ns vs. μs) and not sensitive to temperature fluctuation. Therefore, photodiodes usually work as solar cells and photoconductors. Phototransistors are transistors with the base terminal exposed instead of sending photocurrent into the base. Compared to photodiodes, phototransistors have a greater gain.
The application of TMDCs in photodetectors has been widely demonstrated. Basically, the metal–semiconductor–metal (M-S-M) structured photodetector is made of MoS2 and amorphous silicon with response times of about 0.3 ms and a photoresponsivity of 210 mA/W when irradiated with green light [190]. By using a two-pulse photo-voltage correlation technique, an ultrafast response of 3 ps was recorded in photodetectors constructed with monolayer MoS2 [208]. A phototransistor made from mechanical exfoliated monolayer MoS2 has also exhibited high photo-responsivity of 7.5 mA/W. By improved mobility and contact quality, an ultrasensitive monolayer MoS2 photodetector was also fabricated [209]. The maximum external photo-responsivity was able to attain up to 880 A/W. As shown in Figure 15a, the drain current Ids can increase upon light illumination. This device also showed a uniform increase in photo-responsivity as the illumination wavelength was reduced from 680 to 400 nm (Figure 15b). This suggests that monolayer MoS2 photodetectors can be used for a broad range of wavelengths, between 400 and 650 nm. Phototransistors based on CVD-grown monolayer MoS2 and multilayer WS2 have also been demonstrated [210,211]. Photodetectors with high photo-responsivity (2.5 × 107 A/W) and external quantum efficiency (3168%) were fabricated by multilayer ReS2 [75,78].
As it is a classic structure of TMDC-based photodetectors, the M-S-M configuration has been systematically studied. Schottky barrier modulation between 2D-TMDCs and electrodes is one of the promising methods of improving performance [212,213,214]. Phase transformation of MoS2 (2H-1T) will also reduce the native Schottky barrier and increase photo-responsivity by more than one order of magnitude [215]. Self-assembled doping is another method of improving mobility and photo-responsivity [216]. After all, the photo-responsivity of the TMDC photodetectors (~880 mA/W) [209] is much greater than that of graphene photodetectors (15.7 mA/W) [217], and the general response speed (~3 ps on MoS2) [208] is relatively high (~50 ps on graphene) [218].
The photothermoelectric effect also plays a considerable role in photo-responsivity [219]. In addition to photo-excited electron−hole pairs across the Schottky barriers at the TMDC/electrode interface, the carrier is also excited by light illumination based on temperature difference. The photovoltaic effect interface dominates in the accumulation regime, whereas the hot-carrier-assisted photothermoelectric effect prevails in the depletion regime for the multilayer MoS2 photodetector [220]. Meanwhile, the photocurrent generation in monolayer MoS2 is dominated by the photothermoelectric effect [221]. The photothermoelectric effect has also been studied in other TMDC-based photodetectors [153,222].
By using MoS2 layers of different thicknesses, phototransistors can be tuned to absorb lights at different wavelengths, as shown in Figure 16 [198]. Furthermore, multilayer MoS2 phototransistors have been demonstrated to response spectra range from ultraviolet (UV) to near-infrared (NIR) [223]. Similar bandgap tuning, by varying the number of TMDC layers, has also been applied for WS2 [224].
Heterojunction fabrication can be performed to optimize the performance of 2D-TMDC photodetectors. Highly efficient photocurrent generation has been demonstrated in vertical heterostructures of Gr/TMDCs and Gr/hBN/TMDCs [177,225], where a maximum external quantum efficiency was 55%, internal quantum efficiency was up to 85%, and the response speed was down to 53.6 μs [179]. The hBN/Gr/WSe2 vertical heterostructure photodetectors have a superfast photo-response time of 5.5 ps [180]. Then, a photodetector with a WSe2/Gr/MoS2 heterostructure, which extends the absorption spectral range from 400 to 2400 nm [181].
In addition, atomically 2D-TMDCs still have physical limitations with light absorption, so the use of an absorption layer (e.g., R6G/PbS/perovskite) was applied on 2D-TMDC photodetectors [226,227,228].

3.4. Lasers

In photoluminescence (PL) devices, light emission will occur after absorbing photons at greater energy. The enhancement of PL quantum yield has been demonstrated in TMDC monolayers [16,229,230,231,232], although the yield difference between bulk and monolayer MoS2 was relatively modest. The absolute PL intensity of the WS2 and WSe2 monolayers was 20–40 times greater than that of monolayer MoS2 exfoliated from a natural crystal [22,231]. Compared to MoS2, WS2 and WSe2 are better candidates for light emission devices.
However, the spontaneous emission efficiency of monolayer TMDCs is still low because the non-radiative recombination rate exceeds the spontaneous emission rate [233]. For monolayer MoS2 on SiO2 substrates, the non-radiative decay time is ~70 ps and the spontaneous emission lifetime is ~10 ns, as estimated at room temperature [234,235].
The enhancement of spontaneous emission could be achieved by two major strategies as shown in Figure 17a,b. One is to couple the TMDCs with a low Q-factor (~300) planar photonic crystal (PPC) [236,237]. The other one is to integrate TMDCs with distributed Bragg reflector (DBR) micro-cavities [238,239]. Recently, device performance based on monolayer WSe2 was improved by using the former strategy. The as-fabricated PCCs had Q-factors of about 104. This led to a significant improvement of the Purcell factor, which is crucial for lasing. The optical pumping threshold was as low as 27 nW at 130 K [240].
Besides the Purcell enhancement of spontaneous emission, there are several critical challenges for achieving lasing using 2D TMDCs. For example, the large effective masses of charge carriers in MoS2 result in high densities of states in both valence and conduction bands. MoS2 with 1.8 eV bandgap requires carrier concentrations to push the quasi-Fermi levels into the corresponding bands to achieve population inversion [235]. The transition between the valence band and the conduction band-edge has a strong excitonic feature in such a monolayer TMDC system, including neutral and red-shifted charged excitons, allowing for the long-lived population inversion required, which has been studied in MoS2 [241], WS2 [242], and WSe2 [243].
In addition, the limited gains and the lack of optical confinement and feedback within the monolayer TMDCs will hinder coherent light emission. There have been demonstrations that a strong optical confinement and an enhanced modal gain can be achieved by embedding 2D TMDCs at the interface between a free-standing microsphere and a microdisk [235,244], as shown in Figure 17c,d. The devices exhibited multiple resonant lasing peaks in the wavelength range of ∼575–775 nm. In Figure 17d, the structure of an Si3N4/WS2/HSQ microdisk with a diameter of 3.3 μm has two advantages: enhanced optical mode overlap and material protection. The lasing performance did not decay even after one year, as the sandwich structure protects the monolayer from direct exposure to air [244].
It was noted that the key to the lasing lies in the monolayer of the gain medium, which confines direct-gap excitons to within micro- or even nano-structures. The configuration allows for gain properties via external controls such as electrostatic gating and current injection, enabling optically or electrically pumped operation. These results show that the possibilities of these fabrication methods are scalable or designable, and compatible with integrated electronic circuits technologies. However, the performance of the device might be further improved by exploring surface passivation, doping, and strain engineering.

4. Integrated Circuits

4.1. Amplifiers and Inverters

An amplifier is a basic FET device that utilizes power from the supply to amplify the input signals to greater amplitudes. The operation of an analog small signal amplifier was demonstrated based on monolayer MoS2 [102]. As shown in Figure 18a, the device consisted of two transistors that were integrated on a single monolayer MoS2. The gate of one transistor (the switch) acted to provide input, while the gate of the other transistor (the load) was connected to the central lead and acted as the output/amplifier. The power supply of the amplifier VDD was set to 2 V. The switch and load transistors connected in series are represented in Figure 18b. The switch transistor was first biased at a certain DC voltage to establish a desired drain current, shown as the Q-point (quiescent point) in Figure 18c. When a small sinusoidal AC signal Vin-AC of amplitude ∆Vin/2 was superimposed on the DC signal at the input, the output voltage ∆Vout oscillated synchronously with a phase difference of 180° with respect to Vin-AC. The slope of the red straight line in Figure 18c represented the gain
G = Δ V o u t Δ V i n > 4
of the amplifier at room temperature. By increasing the frequency of the AC signal, the gain was reduced to 1 at 2000 Hz, as shown in Figure 18d.
An important application of amplifiers is logic gate inverters (NOT gates). A NOT gate is designed to convert Logical 0 (low input voltage) to Logical 1 (high output voltage), and vice versa. The first monolayer MoS2 inverter was reported in 2011 [97]. Up to six independently switchable transistors were fabricated on a single monolayer MoS2 flake by lithographically patterning. As shown in Figure 19, an NOR-gate logic operation was also demonstrated, making them suitable for incorporation into digital circuits. A NOR gate is a universal gate that can be built in combinations to form all other logic operations.
On the other hand, radio frequency (RF) amplifiers and a logic inverter can be formed by integrating multiple MoS2 transistors on quartz or flexible substrates. These devices have been demonstrated with an intrinsic gain over 30 and work in the GHz regime [90]. Several reports on amplifiers or inverters based on TMDCs were also reported with a voltage gain ranging from 1.45 to 27 [58,91,97,116,165,245,246,247,248]. Low-power consumption complementary inverters were also fabricated in the sub-nanowatt range [248], which is the most important advantage of CMOS inverters over single FET-type inverters.

4.2. Logic Circuits

Besides amplifier and inverters, more complicated devices of logic circuits have also been investigated. Wang et al. demonstrated complex integrated circuits based on bilayer MoS2. These ICs consist of inverters, logical NAND gates, a static random access memory (SRAM), and a five-stage ring oscillator [249]. It is noted that an NAND gate is another basic logic gate with universal functionality. Any other type of logic gate can then be constructed with a combination of NAND gates. The schematic design and the optical micrograph of an NAND gate and SRAM circuit fabricated on the bilayer MoS2 is shown in Figure 20. The SRAM was constructed from a pair of cross-coupled inverters as shown. This storage cell had two stable states (0 and 1) at the output. The flip-flop cell could be set to Logic State 1 (or 0) by applying a low (or high) voltage to the input.
The five-stage ring oscillator (Figure 21A,B) was constructed to assess the high frequency switching capability and for evaluating MoS2 ultimate compatibility with conventional circuit architecture. The positive feedback loop in the ring oscillator resulted in a statically unstable system, and the voltage at the output of each inverter stage oscillated as a function of time as shown in Figure 21C. The output signal of the ring oscillator can also be measured in terms of its frequency power spectrum. As shown in Figure 21D, the fundamental oscillation frequency was at 1.6 MHz at Vdd = 2 V, corresponding to a propagation delay of 62.5 ns per stage [249].

4.3. Memory Devices

Memory cells are more basic building blocks of integrated digital electronics. In fact, graphene-based nonvolatile memory devices were studied earlier because of the high carrier mobility. However, graphene memory devices typically exhibited very low program/erase (P/E) current ratios up to 5.5 due to the absence of an energy bandgap. To overcome this limitation in graphene-based memory devices, TMDCs with a sufficient bandgap was proposed as a channel material for nonvolatile memory devices. The TMDC FETs showed a high on/off-current ratio and have the potential to enhance the P/E ratio in nonvolatile memory systems. The first nonvolatile memory cell based on TMDCs was demonstrated in 2012 with monolayer MoS2 [101]. The P/E ratio can reach 5 × 103 at 1 V, as shown in Table 6.
Furthermore, nonvolatile memory devices that combine graphene with monolayer MoS2, which reduced the operation voltage (Vds) to 0.05 V, have been demonstrated [252]. As shown in Figure 22a, monolayer MoS2 is used as the semiconducting channel with graphene stripes as the electrodes in FET geometry. A piece of multilayer graphene (MLG), separated by a 6-nm-thick tunneling HfO2 layer from the monolayer MoS2 channel was used as the floating gate. The conductivity of the MoS2 channel was modulated by the voltage VCG applied on the top Cr/Au control gate electrode (CG). When a positive VCG is applied to the device, electrons will tunnel from the MoS2 channel through the HfO2 and accumulate on the MLG floating gate. The application of a negative VCG will deplete the floating gate and reset the device, as shown in Figure 22b. The MLG has a work function of 4.6 eV, which is not sensitive to the number of layers and results in a deep potential well for charge trapping, thus improving the charge retention. Monolayer MoS2 was highly sensitive to the presence of charges in the floating gate MLG, resulting in a 104 difference between memory program and erase states in Figure 22c.
In addition, a new type of the memory device was fabricated with graphene (G) as the FET channel, hBN (B) as the tunnel barrier, and MoS2 (M) as the charge trapping layer (denoted as GBM) [253,254]. This result confirmed that the MoS2 layer could act as an effective charge-trapping layer. These GBM devices also showed different hysteresis characteristics, depending on the thicknesses of MoS2 and hBN. When a thicker MoS2 layer and thinner hBN were employed, unipolar conductance and greater hysteresis were observed due to effective electron tunneling and electric-field screening. In addition, the reversed stacking structure (MBG) was also investigated, with a high on/off current ratio and a large memory window [253]. Due to the time-dependent PL, the MoS2–graphene heterostructure can also function as a rewritable optoelectronic switch or memory, where the persistent state shows almost no relaxation or decay within experimental timescales, indicating near-perfect charge retention [255].
The combination of TMDCs with ferroelectric materials was also used for nonvolatile memory devices as schematically drawn in Figure 23 [101]. FETs based on monolayer or multilayer MoS2 were fabricated on a lead zirconium titanate (Pb(Zr,Ti)O3, PZT) substrates. The PZT substrate can be used as the alternative to the optoelectronic switching, to write and erase the data as shown in Figure 23. When the MoS2 layer was illuminated by visible light, the photo-generated carrier would create an internal electric field to affect the polarization of PZT beneath the MoS2. Only after 5 min illumination, the ON and OFF states became indistinguishable, and data was erased [250].
A type-switching memory device formed with a BP/hBN/MoS2 heterostructure was developed based on the ambipolar property of BP memory devices [256], and the memory device operated under p-type and n-type modes according to the polarity of a gate voltage pulse, but the P/E ratio was only up to 50 [113], which is not comparable to the Gr/MoS2-based devices. These MoS2-based memory devices exhibited high-performance in terms of P/E ratio (105), long endurance, and retention (105 cycles) and have a low operating voltage (0.05 V). It is expected that TMDCs can be used for non-volatile, portable, and power saving memory devices.

5. Summary

In this review, we have critically evaluated the progress of electronic and optoelectronic devices based on 2D TMDCs. Apparently, the application of TMDCs for field effect transistors (FETs) has been very popular. Most efforts have focused on improving the device mobility by various approaches, including encapsulation with dielectric films, gate configurations, managing the contact barriers, and using “defect-free” substrates such as BN nanosheets. The progress in electronic devices has now involved the development of complex integrated circuits (ICs) and logic gates. Many of these are based on the layout design of the circuit on a same piece of 2D TMDC, fully demonstrating the advantage of “large-area” 2D materials for device integration. On the other hand, many advances have been demonstrated for optoelectronic devices, including solar cells, light-emitting diodes (LEDs), photodetectors, and even lasers. Overall, the efficiency of the TMDC-based optoelectronics is still low. Some of the major issues have been the difficulty in p-type and n-type doping on 2D TMDCs. Other issues are due to the atomically thin nature of the materials, which prevent sufficient light absorption, in spite of their high absorption coefficient (~107 m−1 in the visible range) [257]. “Stacked heterostructures” appear to be a potential solution, either by stacking multiple 2D TMDCs or other 2D materials (graphene and BN nanosheets), or stacking on other functional “substrates” such as photonic structures, among others. However, most of the 2D heterostructure devices were fabricated by mechanical exfoliation and transference, which is complicated and inefficient for opto-electro applications. The improvement of 2D heterostructure synthesis on a large scale is critical for developing the 2D materials.
It is noted that TMDCs are among the very few semiconducting 2D materials that are stable for device applications. The semiconducting properties in TMDCs are the important features that supplement the gapless graphene and the electrically insulating hBN for 2D devices. The emergence of TMDCs has enabled meaningful applications of vertically stacked heterostructures for novel electronics and optoelectronics devices with tailorable properties. Overall, the progress on 2D TMDC devices is still in its infancy and requires further development. The biggest issue of the TMDCs is the relatively low performance of their electronic and optoelectronic devices, as compared to those constructed by other state-of-the-art semiconductors. The major advantages of TMDC-based devices are the use of very few materials (atomically thin) and the capability of being fabricated as flexible and wearable devices. TMDCs enable energy/materials saving fabrication of highly portable devices, which meet the desired features of nanotechnology.

Acknowledgments

Yoke Khin Yap acknowledges the support from the National Science Foundation (Award number DMR-1261910).

Author Contributions

M.Y. wrote the manuscript with the supervision from D.Z. and Y.K.Y. D.Z. and Y.K.Y. plan, draft, edit and rewrite the manuscript.

Conflicts of Interest

The authors declare no conflict of interest. The founding sponsors had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript; or in the decision to publish the results.

References

  1. Lundstrom, M. Applied physics. Moore’s law forever? Science 2003, 299, 210–211. [Google Scholar] [CrossRef] [PubMed]
  2. Theis, T.N.; Solomon, P.M. It’s time to reinvent the transistor! Science 2010, 327, 1600–1601. [Google Scholar] [CrossRef] [PubMed]
  3. Morton, J.J.; McCamey, D.R.; Eriksson, M.A.; Lyon, S.A. Embracing the quantum limit in silicon computing. Nature 2011, 479, 345–353. [Google Scholar] [CrossRef] [PubMed]
  4. Ferain, I.; Colinge, C.A.; Colinge, J.P. Multigate transistors as the future of classical metal-oxide-semiconductor field-effect transistors. Nature 2011, 479, 310–316. [Google Scholar] [CrossRef] [PubMed]
  5. Ionescu, A.M.; Riel, H. Tunnel field-effect transistors as energy-efficient electronic switches. Nature 2011, 479, 329–337. [Google Scholar] [CrossRef] [PubMed]
  6. Lee, C.H.; Qin, S.; Savaikar, M.A.; Wang, J.; Hao, B.; Zhang, D.; Banyai, D.; Jaszczak, J.A.; Clark, K.W.; Idrobo, J.C.; et al. Room-Temperature Tunneling Behavior of Boron Nitride Nanotubes Functionalized with Gold Quantum Dots. Adv. Mater. 2013, 25, 4544–4548. [Google Scholar] [CrossRef] [PubMed]
  7. Hao, B.; Asthana, A.; Hazaveh, P.K.; Bergstrom, P.L.; Banyai, D.; Savaikar, M.A.; Jaszczak, J.A.; Yap, Y.K. New Flexible Channels for Room Temperature Tunneling Field Effect Transistors. Sci. Rep. 2016, 6. [Google Scholar] [CrossRef] [PubMed]
  8. Parashar, V.; Durand, C.P.; Hao, B.; Amorim, R.G.; Pandey, R.; Tiwari, B.; Zhang, D.; Liu, Y.; Li, A.P.; Yap, Y.K. Switching Behaviors of Graphene-Boron Nitride Nanotube Heterojunctions. Sci. Rep. 2015, 5, 12238. [Google Scholar] [CrossRef] [PubMed]
  9. Shim, J.; Park, H.-Y.; Kang, D.-H.; Kim, J.-O.; Jo, S.-H.; Park, Y.; Park, J.-H. Electronic and Optoelectronic Devices based on Two-Dimensional Materials: From Fabrication to Application. Adv. Electron. Mater. 2017, 3, 1600364. [Google Scholar] [CrossRef]
  10. Ye, M.; Winslow, D.; Zhang, D.; Pandey, R.; Yap, Y. Recent Advancement on the Optical Properties of Two-Dimensional Molybdenum Disulfide (MoS2) Thin Films. Photonics 2015, 2, 288–307. [Google Scholar] [CrossRef]
  11. Tongay, S.; Zhou, J.; Ataca, C.; Lo, K.; Matthews, T.S.; Li, J.; Grossman, J.C.; Wu, J. Thermally driven crossover from indirect toward direct bandgap in 2D semiconductors: MoSe2 versus MoS2. Nano Lett. 2012, 12, 5576–5580. [Google Scholar] [CrossRef] [PubMed]
  12. Zhang, Y.; Chang, T.R.; Zhou, B.; Cui, Y.T.; Yan, H.; Liu, Z.; Schmitt, F.; Lee, J.; Moore, R.; Chen, Y.; et al. Direct observation of the transition from indirect to direct bandgap in atomically thin epitaxial MoSe2. Nat. Nanotechnol. 2014, 9, 111–115. [Google Scholar] [CrossRef] [PubMed]
  13. Goldberg, A.M.; Beal, A.R.; Lévy, F.A.; Davis, E.A. The low-energy absorption edge in 2H-MoS2 and 2H-MoSe2. Philos. Mag. 1975, 32, 367–378. [Google Scholar] [CrossRef]
  14. Coehoorn, R.; Haas, C.; Dijkstra, J.; Flipse, C.J.; de Groot, R.A.; Wold, A. Electronic structure of MoSe2, MoS2, and WSe2. I. Band-structure calculations and photoelectron spectroscopy. Phys. Rev. B Condens. Matter 1987, 35, 6195–6202. [Google Scholar] [CrossRef] [PubMed]
  15. Ma, Y.; Dai, Y.; Guo, M.; Niu, C.; Lu, J.; Huang, B. Electronic and magnetic properties of perfect, vacancy-doped, and nonmetal adsorbed MoSe2, MoTe2 and WS2 monolayers. Phys. Chem. Chem. Phys. 2011, 13, 15546–15553. [Google Scholar] [CrossRef] [PubMed]
  16. Ruppert, C.; Aslan, O.B.; Heinz, T.F. Optical properties and band gap of single- and few-layer MoTe2 crystals. Nano Lett. 2014, 14, 6231–6236. [Google Scholar] [CrossRef] [PubMed]
  17. Grant, A.J.; Griffiths, T.M.; Pitt, G.D.; Yoffe, A.D. The electrical properties and the magnitude of the indirect gap in the semiconducting transition metal dichalcogenide layer crystals. J. Phys. C Solid State Phys. 1975, 8, L17–L23. [Google Scholar] [CrossRef]
  18. Hind, S.P.; Lee, P.M. KKR calculations of the energy bands in NbSe2, MoS2 and alpha MoTe2. J. Phys. C Solid State Phys. 1980, 13, 349–357. [Google Scholar] [CrossRef]
  19. Ding, Y.; Wang, Y.; Ni, J.; Shi, L.; Shi, S.; Tang, W. First principles study of structural, vibrational and electronic properties of graphene-like MX2 (M = Mo, Nb, W, Ta; X = S, Se, Te) monolayers. Phys. B Condens. Matter 2011, 406, 2254–2260. [Google Scholar] [CrossRef]
  20. Jo, S.; Ubrig, N.; Berger, H.; Kuzmenko, A.B.; Morpurgo, A.F. Mono- and bilayer WS2 light-emitting transistors. Nano Lett. 2014, 14, 2019–2025. [Google Scholar] [CrossRef] [PubMed]
  21. Kam, K.K.; Parklnclon, B.A. Detailed Photocurrent Spectroscopy of the Semiconducting Grouping VI Transition Metal Dichalcogenides. J. Phys. Chem. 1982, 86, 463–467. [Google Scholar] [CrossRef]
  22. Zhao, W.; Ghorannevis, Z.; Chu, L.; Toh, M.; Kloc, C.; Tan, P.-H.; Eda, G. Evolution of Electronic Structure in Atomically Thin sheets of WS2 and WSe2. ACS Nano 2013, 7, 791–797. [Google Scholar] [CrossRef] [PubMed]
  23. Dawson, W.G.; Bullett, D.W. Electronic-Structure and Crystallography of MoTe2 and Wte2. J. Phys. C Solid State Phys. 1987, 20, 6159–6174. [Google Scholar] [CrossRef]
  24. Tongay, S.; Sahin, H.; Ko, C.; Luce, A.; Fan, W.; Liu, K.; Zhou, J.; Huang, Y.S.; Ho, C.H.; Yan, J.; et al. Monolayer behaviour in bulk ReS2 due to electronic and vibrational decoupling. Nat. Commun. 2014, 5, 3252. [Google Scholar] [CrossRef] [PubMed]
  25. Friemelt, K.; Kulikova, L.; Kulyuk, L.; Siminel, A.; Arushanov, E.; Kloc, C.; Bucher, E. Optical and photoelectrical properties of ReS2 single crystals. J. Appl. Phys. 1996, 79, 9268–9272. [Google Scholar] [CrossRef]
  26. Wolverson, D.; Crampin, S.; Kazemi, A.S.; Ilie, A.; Bending, S.J. Raman Spectra of Monolayer, Few-Layer, and Bulk ReSe2: An Anisotropic Layered Semiconductor. ACS Nano 2014, 8, 11154–11164. [Google Scholar] [CrossRef] [PubMed]
  27. Yang, S.; Tongay, S.; Li, Y.; Yue, Q.; Xia, J.B.; Li, S.S.; Li, J.; Wei, S.H. Layer-dependent electrical and optoelectronic responses of ReSe2 nanosheet transistors. Nanoscale 2014, 6, 7226–7231. [Google Scholar] [CrossRef] [PubMed]
  28. Yang, S.; Wang, C.; Sahin, H.; Chen, H.; Li, Y.; Li, S.S.; Suslu, A.; Peeters, F.M.; Liu, Q.; Li, J.; et al. Tuning the optical, magnetic, and electrical properties of ReSe2 by nanoscale strain engineering. Nano Lett. 2015, 15, 1660–1666. [Google Scholar] [CrossRef] [PubMed]
  29. Zhao, H.; Wu, J.B.; Zhong, H.X.; Guo, Q.S.; Wang, X.M.; Xia, F.N.; Yang, L.; Tan, P.H.; Wang, H. Interlayer interactions in anisotropic atomically thin rhenium diselenide. Nano Res. 2015, 8, 3651–3661. [Google Scholar] [CrossRef]
  30. Marzik, J.V.; Kershaw, R.; Dwight, K.; Wold, A. Photoelectronic properties of ReS2 and ReSe2 single crystals. J. Solid State Chem. 1984, 51, 170–175. [Google Scholar] [CrossRef]
  31. Ridley, B.K. The electron-phonon interaction in quasi-two-dimensional semiconductor quantum-well structures. J. Phys. C Solid State Phys. 1982, 15, 5899–5917. [Google Scholar] [CrossRef]
  32. Chen, J.H.; Jang, C.; Xiao, S.; Ishigami, M.; Fuhrer, M.S. Intrinsic and extrinsic performance limits of graphene devices on SiO2. Nat. Nanotechnol. 2008, 3, 206–209. [Google Scholar] [CrossRef] [PubMed]
  33. Fivaz, R.; Mooser, E. Mobility of Charge Carriers in Semiconducting Layer Structures. Phys. Rev. 1967, 163, 743–755. [Google Scholar] [CrossRef]
  34. Ando, T.; Fowler, A.B.; Stern, F. Electronic properties of two-dimensional systems. Rev. Mod. Phys. 1982, 54, 437–672. [Google Scholar] [CrossRef]
  35. Hwang, E.H.; Adam, S.; Sarma, S.D. Carrier transport in two-dimensional graphene layers. Phys. Rev. Lett. 2007, 98, 186806. [Google Scholar] [CrossRef] [PubMed]
  36. Hess, K.; Vogl, P. Remote polar phonon scattering in silicon inversion layers. Solid State Commun. 1979, 30, 797–799. [Google Scholar] [CrossRef]
  37. Nomura, K.; MacDonald, A.H. Quantum transport of massless Dirac fermions. Phys. Rev. Lett. 2007, 98, 076602. [Google Scholar] [CrossRef] [PubMed]
  38. Adam, S.; Hwang, E.H.; Sarma, S.D. Scattering mechanisms and Boltzmann transport in graphene. Phys. E Low-Dimens. Syst. Nanostruct. 2008, 40, 1022–1025. [Google Scholar] [CrossRef]
  39. Paul, T.; Ghatak, S.; Ghosh, A. Percolative switching in transition metal dichalcogenide field-effect transistors at room temperature. Nanotechnology 2016, 27, 125706. [Google Scholar] [CrossRef] [PubMed]
  40. Podzorov, V.; Gershenson, M.E.; Kloc, C.; Zeis, R.; Bucher, E. High-mobility field-effect transistors based on transition metal dichalcogenides. Appl. Phys. Lett. 2004, 84, 3301–3303. [Google Scholar] [CrossRef]
  41. Novoselov, K.S.; Jiang, D.; Schedin, F.; Booth, T.J.; Khotkevich, V.V.; Morozov, S.V.; Geim, A.K. Two-dimensional atomic crystals. Proc. Natl. Acad. Sci. USA 2005, 102, 10451–10453. [Google Scholar] [CrossRef] [PubMed]
  42. Ayari, A.; Cobas, E.; Ogundadegbe, O.; Fuhrer, M.S. Realization and electrical characterization of ultrathin crystals of layered transition-metal dichalcogenides. J. Appl. Phys. 2007, 101, 014507. [Google Scholar] [CrossRef]
  43. Lee, K.; Kim, H.Y.; Lotya, M.; Coleman, J.N.; Kim, G.T.; Duesberg, G.S. Electrical characteristics of molybdenum disulfide flakes produced by liquid exfoliation. Adv. Mater. 2011, 23, 4178–4182. [Google Scholar] [CrossRef] [PubMed]
  44. Zhu, W.J.; Perebeinos, V.; Freitag, M.; Avouris, P. Carrier scattering, mobilities, and electrostatic potential in monolayer, bilayer, and trilayer graphene. Phys. Rev. B 2009, 80, 235402. [Google Scholar] [CrossRef]
  45. Lee, Y.H.; Zhang, X.Q.; Zhang, W.; Chang, M.T.; Lin, C.T.; Chang, K.D.; Yu, Y.C.; Wang, J.T.; Chang, C.S.; Li, L.J.; et al. Synthesis of large-area MoS2 atomic layers with chemical vapor deposition. Adv. Mater. 2012, 24, 2320–2325. [Google Scholar] [CrossRef] [PubMed]
  46. Kim, S.; Konar, A.; Hwang, W.S.; Lee, J.H.; Lee, J.; Yang, J.; Jung, C.; Kim, H.; Yoo, J.B.; Choi, J.Y.; et al. High-mobility and low-power thin-film transistors based on multilayer MoS2 crystals. Nat. Commun. 2012, 3, 1011. [Google Scholar] [CrossRef] [PubMed]
  47. Qiu, H.; Pan, L.J.; Yao, Z.N.; Li, J.J.; Shi, Y.; Wang, X.R. Electrical characterization of back-gated bi-layer MoS2 field-effect transistors and the effect of ambient on their performances. Appl. Phys. Lett. 2012, 100, 123104. [Google Scholar]
  48. Liu, H.; Ye, P.D. MoS2 Dual-Gate MOSFET With Atomic-Layer-Deposited Al2O3 as Top-Gate Dielectric. IEEE Electron. Device Lett. 2012, 33, 546–548. [Google Scholar] [CrossRef]
  49. Das, S.; Chen, H.Y.; Penumatcha, A.V.; Appenzeller, J. High performance multilayer MoS2 transistors with scandium contacts. Nano Lett. 2013, 13, 100–105. [Google Scholar] [CrossRef] [PubMed]
  50. Wu, W.; De, D.; Chang, S.C.; Wang, Y.; Peng, H.; Bao, J.; Pei, S.-S. High mobility and high on/off ratio field-effect transistors based on chemical vapor deposited single-crystal MoS2 grains. Appl. Phys. Lett. 2013, 102, 142106. [Google Scholar] [CrossRef]
  51. Pradhan, N.R.; Rhodes, D.; Zhang, Q.; Talapatra, S.; Terrones, M.; Ajayan, P.M.; Balicas, L. Intrinsic carrier mobility of multi-layered MoS2 field-effect transistors on SiO2. Appl. Phys. Lett. 2013, 102, 123105. [Google Scholar] [CrossRef]
  52. Bao, W.Z.; Cai, X.H.; Kim, D.; Sridhara, K.; Fuhrer, M.S. High mobility ambipolar MoS2 field-effect transistors: Substrate and dielectric effects. Appl. Phys. Lett. 2013, 102, 042104. [Google Scholar] [CrossRef]
  53. Desai, S.B.; Madhvapathy, S.R.; Sachid, A.B.; Llinas, J.P.; Wang, Q.; Ahn, G.H.; Pitner, G.; Kim, M.J.; Bokor, J.; Hu, C.; Wong, H.S.P.; Javey, A. MoS2 transistors with 1-nanometer gate lengths. Science 2016, 354, 99–102. [Google Scholar] [CrossRef] [PubMed]
  54. Larentis, S.; Fallahazad, B.; Tutuc, E. Field-effect transistors and intrinsic mobility in ultra-thin MoSe2 layers. Appl. Phys. Lett. 2012, 101, 223104. [Google Scholar] [CrossRef]
  55. Chamlagain, B.; Li, Q.; Ghimire, N.J.; Chuang, H.J.; Perera, M.M.; Tu, H.; Xu, Y.; Pan, M.; Xiao, D.; Yan, J.; Mandrus, D.; Zhou, Z. Mobility improvement and temperature dependence in MoSe2 field-effect transistors on parylene-C substrate. ACS Nano 2014, 8, 5079–5088. [Google Scholar] [CrossRef] [PubMed]
  56. Pradhan, N.R.; Rhodes, D.; Xin, Y.; Memaran, S.; Bhaskaran, L.; Siddiq, M.; Hill, S.; Ajayan, P.M.; Balicas, L. Ambipolar molybdenum diselenide field-effect transistors: Field-effect and Hall mobilities. ACS Nano 2014, 8, 7923–7929. [Google Scholar] [CrossRef] [PubMed]
  57. Jung, C.; Kim, S.M.; Moon, H.; Han, G.; Kwon, J.; Hong, Y.K.; Omkaram, I.; Yoon, Y.; Kim, S.; Park, J. Highly Crystalline CVD-grown Multilayer MoSe2 Thin Film Transistor for Fast Photodetector. Sci. Rep. 2015, 5, 15313. [Google Scholar] [CrossRef] [PubMed]
  58. Lin, Y.F.; Xu, Y.; Wang, S.T.; Li, S.L.; Yamamoto, M.; Aparecido-Ferreira, A.; Li, W.; Sun, H.; Nakaharai, S.; Jian, W.B.; Ueno, K.; Tsukagoshi, K. Ambipolar MoTe2 transistors and their applications in logic circuits. Adv. Mater. 2014, 26, 3263–3269. [Google Scholar] [CrossRef] [PubMed]
  59. Pradhan, N.R.; Rhodes, D.; Feng, S.; Xin, Y.; Memaran, S.; Moon, B.H.; Terrones, H.; Terrones, M.; Balicas, L. Field-Effect Transistors Based on Few-Layered alpha-MoTe2. ACS Nano 2014, 8, 5911–5920. [Google Scholar] [CrossRef] [PubMed]
  60. Lezama, I.G.; Ubaldini, A.; Longobardi, M.; Giannini, E.; Renner, C.; Kuzmenko, A.B.; Morpurgo, A.F. Surface transport and band gap structure of exfoliated 2H-MoTe2 crystals. 2D Mater. 2014, 1, 021002. [Google Scholar] [CrossRef]
  61. Fathipour, S.; Ma, N.; Hwang, W.S.; Protasenko, V.; Vishwanath, S.; Xing, H.G.; Xu, H.; Jena, D.; Appenzeller, J.; Seabaugh, A. Exfoliated multilayer MoTe2 field-effect transistors. Appl. Phys. Lett. 2014, 105, 192101. [Google Scholar] [CrossRef]
  62. Xu, H.L.; Fathipour, S.; Kinder, E.W.; Seabaugh, A.C.; Fullerton-Shirey, S.K. Reconfigurable Ion Gating of 2H-MoTe2 Field-Effect Transistors Using Poly(ethylene oxide)-CsClO4 Solid Polymer Electrolyte. ACS Nano 2015, 9, 4900–4910. [Google Scholar] [CrossRef] [PubMed]
  63. Yin, L.; Zhan, X.Y.; Xu, K.; Wang, F.; Wang, Z.X.; Huang, Y.; Wang, Q.S.; Jiang, C.; He, J. Ultrahigh sensitive MoTe2 phototransistors driven by carrier tunneling. Appl. Phys. Lett. 2016, 108, 043503. [Google Scholar] [CrossRef]
  64. Octon, T.J.; Nagareddy, V.K.; Russo, S.; Craciun, M.F.; Wright, C.D. Fast High-Responsivity Few-Layer MoTe2 Photodetectors. Adv. Opt. Mater. 2016, 4, 1750–1754. [Google Scholar] [CrossRef]
  65. Sik Hwang, W.; Remskar, M.; Yan, R.; Protasenko, V.; Tahy, K.; Doo Chae, S.; Zhao, P.; Konar, A.; Xing, H.; Seabaugh, A.; et al. Transistors with chemically synthesized layered semiconductor WS2 exhibiting 105 room temperature modulation and ambipolar behavior. Appl. Phys. Lett. 2012, 101, 013107. [Google Scholar] [CrossRef]
  66. Braga, D.; Gutierrez Lezama, I.; Berger, H.; Morpurgo, A.F. Quantitative determination of the band gap of WS2 with ambipolar ionic liquid-gated transistors. Nano Lett. 2012, 12, 5218–5223. [Google Scholar] [CrossRef] [PubMed]
  67. Liu, X.; Hu, J.; Yue, C.; Della Fera, N.; Ling, Y.; Mao, Z.; Wei, J. High performance field-effect transistor based on multilayer tungsten disulfide. ACS Nano 2014, 8, 10396–10402. [Google Scholar] [CrossRef] [PubMed]
  68. Kumar, J.; Kuroda, M.A.; Bellus, M.Z.; Han, S.J.; Chiu, H.Y. Full-range electrical characteristics of WS2 transistors. Appl. Phys. Lett. 2015, 106, 123508. [Google Scholar] [CrossRef]
  69. Chuang, H.J.; Tan, X.; Ghimire, N.J.; Perera, M.M.; Chamlagain, B.; Cheng, M.M.; Yan, J.; Mandrus, D.; Tomanek, D.; Zhou, Z. High mobility WSe2 p- and n-type field-effect transistors contacted by highly doped graphene for low-resistance contacts. Nano Lett. 2014, 14, 3594–3601. [Google Scholar] [CrossRef] [PubMed]
  70. Zhou, H.; Wang, C.; Shaw, J.C.; Cheng, R.; Chen, Y.; Huang, X.; Liu, Y.; Weiss, N.O.; Lin, Z.; Huang, Y.; et al. Large area growth and electrical properties of p-type WSe2 atomic layers. Nano Lett. 2015, 15, 709–713. [Google Scholar] [CrossRef] [PubMed]
  71. Pradhan, N.R.; Rhodes, D.; Memaran, S.; Poumirol, J.M.; Smirnov, D.; Talapatra, S.; Feng, S.; Perea-Lopez, N.; Elias, A.L.; Terrones, M.; et al. Hall and field-effect mobilities in few layered p-WSe(2) field-effect transistors. Sci. Rep. 2015, 5, 8979. [Google Scholar] [CrossRef] [PubMed]
  72. Campbell, P.M.; Tarasov, A.; Joiner, C.A.; Tsai, M.Y.; Pavlidis, G.; Graham, S.; Ready, W.J.; Vogel, E.M. Field-effect transistors based on wafer-scale, highly uniform few-layer p-type WSe2. Nanoscale 2016, 8, 2268–2276. [Google Scholar] [CrossRef] [PubMed]
  73. Liu, B.; Ma, Y.; Zhang, A.; Chen, L.; Abbas, A.N.; Liu, Y.; Shen, C.; Wan, H.; Zhou, C. High-Performance WSe2 Field-Effect Transistors via Controlled Formation of In-Plane Heterojunctions. ACS Nano 2016, 10, 5153–5160. [Google Scholar] [CrossRef] [PubMed]
  74. Corbet, C.M.; McClellan, C.; Rai, A.; Sonde, S.S.; Tutuc, E.; Banerjee, S.K. Field effect transistors with current saturation and voltage gain in ultrathin ReS2. ACS Nano 2015, 9, 363–370. [Google Scholar] [CrossRef] [PubMed]
  75. Zhang, E.; Jin, Y.; Yuan, X.; Wang, W.; Zhang, C.; Tang, L.; Liu, S.; Zhou, P.; Hu, W.; Xiu, F. ReS2-Based Field-Effect Transistors and Photodetectors. Adv. Funct. Mater. 2015, 25, 4076–4082. [Google Scholar] [CrossRef]
  76. Liu, E.; Fu, Y.; Wang, Y.; Feng, Y.; Liu, H.; Wan, X.; Zhou, W.; Wang, B.; Shao, L.; Ho, C.H.; et al. Integrated digital inverters based on two-dimensional anisotropic ReS2 field-effect transistors. Nat. Commun. 2015, 6, 6991. [Google Scholar] [CrossRef] [PubMed]
  77. Xu, K.; Deng, H.X.; Wang, Z.; Huang, Y.; Wang, F.; Li, S.S.; Luo, J.W.; He, J. Sulfur vacancy activated field effect transistors based on ReS2 nanosheets. Nanoscale 2015, 7, 15757–15762. [Google Scholar] [CrossRef] [PubMed]
  78. Shim, J.; Oh, A.; Kang, D.H.; Oh, S.; Jang, S.K.; Jeon, J.; Jeon, M.H.; Kim, M.; Choi, C.; Lee, J.; et al. High-Performance 2D Rhenium Disulfide (ReS2) Transistors and Photodetectors by Oxygen Plasma Treatment. Adv. Mater. 2016, 28, 6985–6992. [Google Scholar] [CrossRef] [PubMed]
  79. Keyshar, K.; Gong, Y.; Ye, G.; Brunetto, G.; Zhou, W.; Cole, D.P.; Hackenberg, K.; He, Y.; Machado, L.; Kabbani, M.; et al. Chemical Vapor Deposition of Monolayer Rhenium Disulfide (ReS2). Adv. Mater. 2015, 27, 4640–4648. [Google Scholar] [CrossRef] [PubMed]
  80. Corbet, C.M.; Sonde, S.S.; Tutuc, E.; Banerjee, S.K. Improved contact resistance in ReSe2 thin film field-effect transistors. Appl. Phys. Lett. 2016, 108, 162104. [Google Scholar] [CrossRef]
  81. Hafeez, M.; Gan, L.; Li, H.; Ma, Y.; Zhai, T. Chemical Vapor Deposition Synthesis of Ultrathin Hexagonal ReSe2 Flakes for Anisotropic Raman Property and Optoelectronic Application. Adv. Mater. 2016, 28, 8296–8301. [Google Scholar] [CrossRef] [PubMed]
  82. Radisavljevic, B.; Radenovic, A.; Brivio, J.; Giacometti, V.; Kis, A. Single-layer MoS2 transistors. Nat. Nanotechnol. 2011, 6, 147–150. [Google Scholar] [CrossRef] [PubMed]
  83. Ghatak, S.; Pal, A.N.; Ghosh, A. Nature of Electronic States in Atomically Thin MoS2 Field-Effect Transistors. ACS Nano 2011, 5, 7707–7712. [Google Scholar] [CrossRef] [PubMed]
  84. Van der Zande, A.M.; Huang, P.Y.; Chenet, D.A.; Berkelbach, T.C.; You, Y.; Lee, G.H.; Heinz, T.F.; Reichman, D.R.; Muller, D.A.; Hone, J.C. Grains and grain boundaries in highly crystalline monolayer molybdenum disulphide. Nat. Mater. 2013, 12, 554–561. [Google Scholar] [CrossRef] [PubMed]
  85. Yazyev, O.V.; Louie, S.G. Electronic transport in polycrystalline graphene. Nat. Mater. 2010, 9, 806–809. [Google Scholar] [CrossRef] [PubMed]
  86. Hwang, S.W.; Remskar, M.; Yan, R.; Kosel, T.; Park, J.K.; Cho, B.J.; Haensch, W.; Xing, H.; Seabaugh, A.; Jena, D. Comparative study of chemically synthesized and exfoliated multilayer MoS2 field-effect transistors. Appl. Phys. Lett. 2013, 102, 043116. [Google Scholar] [CrossRef]
  87. Walia, S.; Balendhran, S.; Wang, Y.C.; Ab Kadir, R.; Zoolfakar, A.S.; Atkin, P.; Ou, J.Z.; Sriram, S.; Kalantar-zadeh, K.; Bhaskaran, M. Characterization of metal contacts for two-dimensional MoS2 nanoflakes. Appl. Phys. Lett. 2013, 103, 232105. [Google Scholar] [CrossRef]
  88. Strait, J.H.; Nene, P.; Rana, F. High intrinsic mobility and ultrafast carrier dynamics in multilayer metal-dichalcogenideMoS2. Phys. Rev. B 2014, 90, 245402. [Google Scholar] [CrossRef]
  89. Das, S.; Appenzeller, J. Where does the current flow in two-dimensional layered systems? Nano Lett. 2013, 13, 3396–3402. [Google Scholar] [CrossRef] [PubMed]
  90. Cheng, R.; Jiang, S.; Chen, Y.; Liu, Y.; Weiss, N.; Cheng, H.C.; Wu, H.; Huang, Y.; Duan, X. Few-layer molybdenum disulfide transistors and circuits for high-speed flexible electronics. Nat. Commun. 2014, 5, 5143. [Google Scholar] [CrossRef] [PubMed]
  91. Wang, H.; Yu, L.; Lee, Y.H.; Fang, W.; Hsu, A.; Herring, P.; Chin, M.; Dubey, M.; Li, L.J.; Kong, J.; et al. Large-scale 2D electronics based on single-layer MoS2 grown by chemical vapor deposition. IEDM Tech. Dig. 2012, 4.6.1–4.6.4. [Google Scholar] [CrossRef]
  92. Das, S.R.; Kwon, J.; Prakash, A.; Delker, C.J.; Das, S.; Janes, D.B. Low-frequency noise in MoSe2 field effect transistors. Appl. Phys. Lett. 2015, 106, 083507. [Google Scholar] [CrossRef]
  93. Kwon, H.-J.; Kang, H.; Jang, J.; Kim, S.; Grigoropoulos, C.P. Analysis of flicker noise in two-dimensional multilayer MoS2 transistors. Appl. Phys. Lett. 2014, 104, 083110. [Google Scholar] [CrossRef]
  94. Na, J.; Joo, M.K.; Shin, M.; Huh, J.; Kim, J.S.; Piao, M.; Jin, J.E.; Jang, H.K.; Choi, H.J.; Shim, J.H.; et al. Low-frequency noise in multilayer MoS2 field-effect transistors: the effect of high-k passivation. Nanoscale 2014, 6, 433–441. [Google Scholar] [CrossRef] [PubMed]
  95. Liu, L.; Kumar, S.B.; Ouyang, Y.; Guo, J. Performance Limits of Monolayer Transition Metal Dichalcogenide Transistors. IEEE Trans. Electron. Devices 2011, 58, 3042–3047. [Google Scholar] [CrossRef]
  96. Yoon, Y.; Ganapathi, K.; Salahuddin, S. How good can monolayer MoS2 transistors be? Nano Lett. 2011, 11, 3768–3773. [Google Scholar] [CrossRef] [PubMed]
  97. Radisavljevic, B.; Whitwick, M.B.; Kis, A. Integrated Circuits and Logic Operations Based on Single-Layer MoS2. ACS Nano 2011, 5, 9934–9938. [Google Scholar] [CrossRef] [PubMed]
  98. Zhan, Y.; Liu, Z.; Najmaei, S.; Ajayan, P.M.; Lou, J. Large-area vapor-phase growth and characterization of MoS2 atomic layers on a SiO2 substrate. Small 2012, 8, 966–971. [Google Scholar] [CrossRef] [PubMed]
  99. Late, D.J.; Liu, B.; Matte, H.S.; Dravid, V.P.; Rao, C.N. Hysteresis in single-layer MoS2 field effect transistors. ACS Nano 2012, 6, 5635–5641. [Google Scholar] [CrossRef] [PubMed]
  100. Lin, M.-W.; Liu, L.Z.; Lan, Q.; Tan, X.B.; Dhindsa, K.S.; Zeng, P.; Naik, V.M.; Cheng, M.M.C.; Zhou, Z.X. Mobility enhancement and highly efficient gating of monolayer MoS2transistors with polymer electrolyte. J. Phys. D Appl. Phys. 2012, 45, 345102. [Google Scholar] [CrossRef]
  101. Lee, H.S.; Min, S.W.; Park, M.K.; Lee, Y.T.; Jeon, P.J.; Kim, J.H.; Ryu, S.; Im, S. MoS2 nanosheets for top-gate nonvolatile memory transistor channel. Small 2012, 8, 3111–3115. [Google Scholar] [CrossRef] [PubMed]
  102. Radisavljevic, B.; Whitwick, M.B.; Kis, A. Small-signal amplifier based on single-layer MoS2. Appl. Phys. Lett. 2012, 101, 043103. [Google Scholar] [CrossRef]
  103. Lembke, D.; Kis, A. Breakdown of high-performance monolayer MoS2 transistors. ACS Nano 2012, 6, 10070–10075. [Google Scholar] [CrossRef] [PubMed]
  104. Baugher, B.W.; Churchill, H.O.; Yang, Y.; Jarillo-Herrero, P. Intrinsic electronic transport properties of high-quality monolayer and bilayer MoS2. Nano Lett. 2013, 13, 4212–4216. [Google Scholar] [CrossRef] [PubMed]
  105. Park, W.; Baik, J.; Kim, T.Y.; Cho, K.; Hong, W.K.; Shin, H.J.; Lee, T. Photoelectron spectroscopic imaging and device applications of large-area patternable single-layer MoS2 synthesized by chemical vapor deposition. ACS Nano 2014, 8, 4961–4968. [Google Scholar] [CrossRef] [PubMed]
  106. Sanne, A.; Ghosh, R.; Rai, A.; Movva, H.C.P.; Sharma, A.; Rao, R.; Mathew, L.; Banerjee, S.K. Top-gated chemical vapor deposited MoS2 field-effect transistors on Si3N4 substrates. Appl. Phys. Lett. 2015, 106, 062101. [Google Scholar] [CrossRef]
  107. Shao, P.-Z.; Zhao, H.M.; Cao, H.W.; Wang, X.F.; Pang, Y.; Li, Y.X.; Deng, N.Q.; Zhang, J.; Zhang, G.Y.; Yang, Y.; et al. Enhancement of carrier mobility in MoS2 field effect transistors by a SiO2 protective layer. Appl. Phys. Lett. 2016, 108, 203105. [Google Scholar] [CrossRef]
  108. Amani, M.; Chin, M.L.; Birdwell, A.G.; O’Regan, T.P.; Najmaei, S.; Liu, Z.; Ajayan, P.M.; Lou, J.; Dubey, M. Electrical performance of monolayer MoS2 field-effect transistors prepared by chemical vapor deposition. Appl. Phys. Lett. 2013, 102, 193107. [Google Scholar] [CrossRef]
  109. Lu, X.; Utama, M.I.; Lin, J.; Gong, X.; Zhang, J.; Zhao, Y.; Pantelides, S.T.; Wang, J.; Dong, Z.; Liu, Z.; et al. Large-area synthesis of monolayer and few-layer MoSe2 films on SiO2 substrates. Nano Lett. 2014, 14, 2419–2425. [Google Scholar] [CrossRef] [PubMed]
  110. Wang, X.; Gong, Y.; Shi, G.; Chow, W.L.; Keyshar, K.; Ye, G.; Vajtai, R.; Lou, J.; Liu, Z.; Ringe, E.; et al. Chemical vapor deposition growth of crystalline monolayer MoSe2. ACS Nano 2014, 8, 5125–5131. [Google Scholar] [CrossRef] [PubMed]
  111. Chang, Y.H.; Zhang, W.; Zhu, Y.; Han, Y.; Pu, J.; Chang, J.K.; Hsu, W.T.; Huang, J.K.; Hsu, C.L.; Chiu, M.H.; et al. Monolayer MoSe2 grown by chemical vapor deposition for fast photodetection. ACS Nano 2014, 8, 8582–8590. [Google Scholar] [CrossRef] [PubMed]
  112. Ovchinnikov, D.; Allain, A.; Huang, Y.S.; Dumcenco, D.; Kis, A. Electrical transport properties of single-layer WS2. ACS Nano 2014, 8, 8174–8181. [Google Scholar] [CrossRef] [PubMed]
  113. Xu, Z.Q.; Zhang, Y.; Lin, S.; Zheng, C.; Zhong, Y.L.; Xia, X.; Li, Z.; Sophia, P.J.; Fuhrer, M.S.; Cheng, Y.B.; et al. Synthesis and Transfer of Large-Area Monolayer WS2 Crystals: Moving Toward the Recyclable Use of Sapphire Substrates. ACS Nano 2015, 9, 6178–6187. [Google Scholar] [CrossRef] [PubMed]
  114. Hanbicki, A.T.; Currie, M.; Kioseoglou, G.; Friedman, A.L.; Jonker, B.T. Measurement of high exciton binding energy in the monolayer transition-metal dichalcogenides WS2 and WSe2. Solid State Commun. 2015, 203, 16–20. [Google Scholar] [CrossRef]
  115. Fang, H.; Chuang, S.; Chang, T.C.; Takei, K.; Takahashi, T.; Javey, A. High-performance single layered WSe(2) p-FETs with chemically doped contacts. Nano Lett. 2012, 12, 3788–3792. [Google Scholar] [CrossRef] [PubMed]
  116. Huang, J.K.; Pu, J.; Hsu, C.L.; Chiu, M.H.; Juang, Z.Y.; Chang, Y.H.; Chang, W.H.; Iwasa, Y.; Takenobu, T.; Li, L.J. Large-Area Synthesis of Highly Crystalline WSe2 Monolayers and Device Applications. ACS Nano 2014, 8, 923–930. [Google Scholar] [CrossRef] [PubMed]
  117. Allain, A.; Kis, A. Electron and hole mobilities in single-layer WSe2. ACS Nano 2014, 8, 7180–7185. [Google Scholar] [CrossRef] [PubMed]
  118. Kawamura, T.; Das Sarma, S. Temperature dependence of the low-temperature mobility in ultrapureAlxGa1−xAs/GaAs heterojunctions: Acoustic-phonon scattering. Phys. Rev. B 1990, 42, 3725–3728. [Google Scholar] [CrossRef]
  119. Hwang, E.H.; Das Sarma, S. Limit to two-dimensional mobility in modulation-doped GaAs quantum structures: How to achieve a mobility of 100 million. Phys. Rev. B 2008, 77, 235437. [Google Scholar] [CrossRef]
  120. Kaasbjerg, K.; Thygesen, K.S.; Jacobsen, K.W. Phonon-limited mobility in n-type single-layer MoS2 from first principles. Phys. Rev. B 2012, 85, 115317. [Google Scholar] [CrossRef]
  121. Zhang, W.X.; Huang, Z.S.; Zhang, W.L.; Li, Y.R. Two-dimensional semiconductors with possible high room temperature mobility. Nano Res. 2014, 7, 1731–1737. [Google Scholar] [CrossRef]
  122. Jena, D.; Konar, A. Enhancement of carrier mobility in semiconductor nanostructures by dielectric engineering. Phys. Rev. Lett. 2007, 98, 136805. [Google Scholar] [CrossRef] [PubMed]
  123. Chen, F.; Xia, J.; Ferry, D.K.; Tao, N. Dielectric screening enhanced performance in graphene FET. Nano Lett. 2009, 9, 2571–2574. [Google Scholar] [CrossRef] [PubMed]
  124. Konar, A.; Fang, T.; Jena, D. Effect of high-κgate dielectrics on charge transport in graphene-based field effect transistors. Phys. Rev. B 2010, 82, 115452. [Google Scholar] [CrossRef]
  125. Newaz, A.K.; Puzyrev, Y.S.; Wang, B.; Pantelides, S.T.; Bolotin, K.I. Probing charge scattering mechanisms in suspended graphene by varying its dielectric environment. Nat. Commun. 2012, 3, 734. [Google Scholar] [CrossRef] [PubMed]
  126. Radisavljevic, B.; Kis, A. Mobility engineering and a metal-insulator transition in monolayer MoS(2). Nat. Mater. 2013, 12, 815–820. [Google Scholar] [CrossRef] [PubMed]
  127. Sakaki, H.; Noda, T.; Hirakawa, K.; Tanaka, M.; Matsusue, T. Interface roughness scattering in GaAs/AlAs quantum wells. Appl. Phys. Lett. 1987, 51, 1934–1936. [Google Scholar] [CrossRef]
  128. Popov, I.; Seifert, G.; Tomanek, D. Designing electrical contacts to MoS2 monolayers: A computational study. Phys. Rev. Lett. 2012, 108, 156802. [Google Scholar] [CrossRef] [PubMed]
  129. Kang, J.H.; Liu, W.; Sarkar, D.; Jena, D.; Banerjee, K. Computational Study of Metal Contacts to Monolayer Transition-Metal Dichalcogenide Semiconductors. Phys. Rev. X 2014, 4, 031005. [Google Scholar] [CrossRef]
  130. Wang, Y.; Yang, R.X.; Quhe, R.; Zhong, H.; Cong, L.; Ye, M.; Ni, Z.; Song, Z.; Yang, J.; Shi, J.; et al. Does p-type ohmic contact exist in WSe2-metal interfaces? Nanoscale 2016, 8, 1179–1191. [Google Scholar] [CrossRef] [PubMed]
  131. Zhong, H.; Quhe, R.; Wang, Y.; Ni, Z.; Ye, M.; Song, Z.; Pan, Y.; Yang, J.; Yang, L.; Lei, M.; et al. Interfacial Properties of Monolayer and Bilayer MoS2 Contacts with Metals: Beyond the Energy Band Calculations. Sci. Rep. 2016, 6, 21786. [Google Scholar] [CrossRef] [PubMed]
  132. Gong, C.; Huang, C.; Miller, J.; Cheng, L.; Hao, Y.; Cobden, D.; Kim, J.; Ruoff, R.S.; Wallace, R.M.; Cho, K.; et al. Metal contacts on physical vapor deposited monolayer MoS2. ACS Nano 2013, 7, 11350–11357. [Google Scholar] [CrossRef] [PubMed]
  133. Leong, W.S.; Luo, X.; Li, Y.; Khoo, K.H.; Quek, S.Y.; Thong, J.T. Low resistance metal contacts to MoS2 devices with nickel-etched-graphene electrodes. ACS Nano 2015, 9, 869–877. [Google Scholar] [CrossRef] [PubMed]
  134. Farmanbar, M.; Brocks, G. Controlling the Schottky barrier at MoS2/metal contacts by inserting a BN monolayer. Phys. Rev. B 2015, 91, 161304. [Google Scholar] [CrossRef]
  135. Chuang, H.J.; Chamlagain, B.; Koehler, M.; Perera, M.M.; Yan, J.; Mandrus, D.; Tomanek, D.; Zhou, Z. Low-Resistance 2D/2D Ohmic Contacts: A Universal Approach to High-Performance WSe2, MoS2, and MoSe2 Transistors. Nano Lett. 2016, 16, 1896–1902. [Google Scholar] [CrossRef] [PubMed]
  136. Farmanbar, M.; Brocks, G. Ohmic Contacts to 2D Semiconductors through van der Waals Bonding. Adv. Electron. Mater. 2016, 2, 1500405. [Google Scholar] [CrossRef]
  137. Chen, J.R.; Odenthal, P.M.; Swartz, A.G.; Floyd, G.C.; Wen, H.; Luo, K.Y.; Kawakami, R.K. Control of Schottky barriers in single layer MoS2 transistors with ferromagnetic contacts. Nano Lett. 2013, 13, 3106–3110. [Google Scholar] [CrossRef] [PubMed]
  138. Dankert, A.; Langouche, L.; Kamalakar, M.V.; Dash, S.P. High-performance molybdenum disulfide field-effect transistors with spin tunnel contacts. ACS Nano 2014, 8, 476–482. [Google Scholar] [CrossRef] [PubMed]
  139. Kappera, R.; Voiry, D.; Yalcin, S.E.; Branch, B.; Gupta, G.; Mohite, A.D.; Chhowalla, M. Phase-engineered low-resistance contacts for ultrathin MoS2 transistors. Nat. Mater. 2014, 13, 1128–1134. [Google Scholar] [CrossRef] [PubMed]
  140. Kappera, R.; Voiry, D.; Yalcin, S.E.; Jen, W.; Acerce, M.; Torrel, S.; Branch, B.; Lei, S.D.; Chen, W.B.; Najmaei, S.; et al. Metallic 1T phase source/drain electrodes for field effect transistors from chemical vapor deposited MoS2. APL Mater. 2014, 2, 092516. [Google Scholar] [CrossRef]
  141. Cho, S.; Kim, S.; Kim, J.H.; Zhao, J.; Seok, J.; Keum, D.H.; Baik, J.; Choe, D.H.; Chang, K.J.; Suenaga, K.; et al. DEVICE TECHNOLOGY. Phase patterning for ohmic homojunction contact in MoTe2. Science 2015, 349, 625–628. [Google Scholar] [CrossRef] [PubMed]
  142. Ma, Y.; Liu, B.; Zhang, A.; Chen, L.; Fathi, M.; Shen, C.; Abbas, A.N.; Ge, M.; Mecklenburg, M.; Zhou, C. Reversible Semiconducting-to-Metallic Phase Transition in Chemical Vapor Deposition Grown Monolayer WSe2 and Applications for Devices. ACS Nano 2015, 9, 7383–7391. [Google Scholar] [CrossRef] [PubMed]
  143. Sangwan, V.K.; Arnold, H.N.; Jariwala, D.; Marks, T.J.; Lauhon, L.J.; Hersam, M.C. Low-frequency electronic noise in single-layer MoS2 transistors. Nano Lett. 2013, 13, 4351–4355. [Google Scholar] [CrossRef] [PubMed]
  144. Xie, X.; Sarkar, D.; Liu, W.; Kang, J.; Marinov, O.; Deen, M.J.; Banerjee, K. Low-frequency noise in bilayer MoS2 transistor. ACS Nano 2014, 8, 5633–5640. [Google Scholar] [CrossRef] [PubMed]
  145. Sharma, D.; Amani, M.; Motayed, A.; Shah, P.B.; Birdwell, A.G.; Najmaei, S.; Ajayan, P.M.; Lou, J.; Dubey, M.; Li, Q.; et al. Electrical transport and low-frequency noise in chemical vapor deposited single-layer MoS2 devices. Nanotechnology 2014, 25, 155702. [Google Scholar] [CrossRef] [PubMed]
  146. Ghatak, S.; Mukherjee, S.; Jain, M.; Sarma, D.D.; Ghosh, A. Microscopic origin of low frequency noise in MoS2 field-effect transistors. APL Mater. 2014, 2, 092515. [Google Scholar] [CrossRef]
  147. Lin, Y.F.; Xu, Y.; Lin, C.Y.; Suen, Y.W.; Yamamoto, M.; Nakaharai, S.; Ueno, K.; Tsukagoshi, K. Origin of Noise in Layered MoTe(2) Transistors and its Possible Use for Environmental Sensors. Adv. Mater. 2015, 27, 6612–6619. [Google Scholar] [CrossRef] [PubMed]
  148. Renteria, J.; Samnakay, R.; Rumyantsev, S.L.; Jiang, C.; Goli, P.; Shur, M.S.; Balandin, A.A. Low-frequency 1/f noise in MoS2 transistors: Relative contributions of the channel and contacts. Appl. Phys. Lett. 2014, 104, 153104. [Google Scholar] [CrossRef]
  149. Zhang, M.; Wu, J.; Zhu, Y.; Dumcenco, D.O.; Hong, J.; Mao, N.; Deng, S.; Chen, Y.; Yang, Y.; Jin, C.; et al. Two-dimensional molybdenum tungsten diselenide alloys: Photoluminescence, Raman scattering, and electrical transport. ACS Nano 2014, 8, 7130–7137. [Google Scholar] [CrossRef] [PubMed]
  150. Zhang, Y.; Ye, J.; Matsuhashi, Y.; Iwasa, Y. Ambipolar MoS2 thin flake transistors. Nano Lett. 2012, 12, 1136–1140. [Google Scholar] [CrossRef] [PubMed]
  151. Zhang, Y.J.; Ye, J.T.; Yomogida, Y.; Takenobu, T.; Iwasa, Y. Formation of a stable p-n junction in a liquid-gated MoS2 ambipolar transistor. Nano Lett. 2013, 13, 3023–3028. [Google Scholar] [CrossRef] [PubMed]
  152. Shi, W.; Ye, J.; Zhang, Y.; Suzuki, R.; Yoshida, M.; Miyazaki, J.; Inoue, N.; Saito, Y.; Iwasa, Y. Superconductivity Series in Transition Metal Dichalcogenides by Ionic Gating. Sci. Rep. 2015, 5, 12534. [Google Scholar] [CrossRef] [PubMed]
  153. Groenendijk, D.J.; Buscema, M.; Steele, G.A.; Michaelis de Vasconcellos, S.; Bratschitsch, R.; van der Zant, H.S.; Castellanos-Gomez, A. Photovoltaic and photothermoelectric effect in a double-gated WSe2 device. Nano Lett. 2014, 14, 5846–5852. [Google Scholar] [CrossRef] [PubMed]
  154. Fontana, M.; Deppe, T.; Boyd, A.K.; Rinzan, M.; Liu, A.Y.; Paranjape, M.; Barbara, P. Electron-hole transport and photovoltaic effect in gated MoS2 Schottky junctions. Sci. Rep. 2013, 3, 1634. [Google Scholar] [CrossRef] [PubMed]
  155. Das, S.; Appenzeller, J. WSe2 field effect transistors with enhanced ambipolar characteristics. Appl. Phys. Lett. 2013, 103, 103501. [Google Scholar] [CrossRef]
  156. Fang, H.; Tosun, M.; Seol, G.; Chang, T.C.; Takei, K.; Guo, J.; Javey, A. Degenerate n-doping of few-layer transition metal dichalcogenides by potassium. Nano Lett. 2013, 13, 1991–1995. [Google Scholar] [CrossRef] [PubMed]
  157. Pospischil, A.; Furchi, M.M.; Mueller, T. Solar-energy conversion and light emission in an atomic monolayer p-n diode. Nat. Nanotechnol. 2014, 9, 257–261. [Google Scholar] [CrossRef] [PubMed]
  158. Baugher, B.W.; Churchill, H.O.; Yang, Y.; Jarillo-Herrero, P. Optoelectronic devices based on electrically tunable p-n diodes in a monolayer dichalcogenide. Nat. Nanotechnol. 2014, 9, 262–267. [Google Scholar] [CrossRef] [PubMed]
  159. Ross, J.S.; Klement, P.; Jones, A.M.; Ghimire, N.J.; Yan, J.; Mandrus, D.G.; Taniguchi, T.; Watanabe, K.; Kitamura, K.; Yao, W.; et al. Electrically tunable excitonic light-emitting diodes based on monolayer WSe2 p-n junctions. Nat. Nanotechnol. 2014, 9, 268–272. [Google Scholar] [CrossRef] [PubMed]
  160. Nakaharai, S.; Yamamoto, M.; Ueno, K.; Lin, Y.F.; Li, S.L.; Tsukagoshi, K. Electrostatically Reversible Polarity of Ambipolar alpha-MoTe2 Transistors. ACS Nano 2015, 9, 5976–5983. [Google Scholar] [CrossRef] [PubMed]
  161. Geim, A.K.; Grigorieva, I.V. Van der Waals heterostructures. Nature 2013, 499, 419–425. [Google Scholar] [CrossRef] [PubMed]
  162. Novoselov, K.S.; Mishchenko, A.; Carvalho, A.; Castro Neto, A.H. 2D materials and van der Waals heterostructures. Science 2016, 353, aac9439. [Google Scholar] [CrossRef] [PubMed]
  163. Yang, H.; Heo, J.; Park, S.; Song, H.J.; Seo, D.H.; Byun, K.E.; Kim, P.; Yoo, I.; Chung, H.J.; Kim, K. Graphene barristor, a triode device with a gate-controlled Schottky barrier. Science 2012, 336, 1140–1143. [Google Scholar] [CrossRef] [PubMed]
  164. Britnell, L.; Gorbachev, R.V.; Jalil, R.; Belle, B.D.; Schedin, F.; Mishchenko, A.; Georgiou, T.; Katsnelson, M.I.; Eaves, L.; Morozov, S.V.; et al. Field-effect tunneling transistor based on vertical graphene heterostructures. Science 2012, 335, 947–950. [Google Scholar] [CrossRef] [PubMed]
  165. Yu, W.J.; Li, Z.; Zhou, H.; Chen, Y.; Wang, Y.; Huang, Y.; Duan, X. Vertically stacked multi-heterostructures of layered materials for logic transistors and complementary inverters. Nat. Mater. 2013, 12, 246–252. [Google Scholar] [CrossRef] [PubMed]
  166. Georgiou, T.; Jalil, R.; Belle, B.D.; Britnell, L.; Gorbachev, R.V.; Morozov, S.V.; Kim, Y.J.; Gholinia, A.; Haigh, S.J.; Makarovsky, O.; et al. Vertical field-effect transistor based on graphene-WS2 heterostructures for flexible and transparent electronics. Nat. Nanotechnol. 2013, 8, 100–103. [Google Scholar] [CrossRef] [PubMed]
  167. Sata, Y.; Moriya, R.; Morikawa, S.; Yabuki, N.; Masubuchi, S.; Machida, T. Electric field modulation of Schottky barrier height in graphene/MoSe2 van der Waals heterointerface. Appl. Phys. Lett. 2015, 107, 023109. [Google Scholar] [CrossRef]
  168. Choi, Y.; Kang, J.; Jariwala, D.; Kang, M.S.; Marks, T.J.; Hersam, M.C.; Cho, J.H. Low-Voltage Complementary Electronics from Ion-Gel-Gated Vertical Van der Waals Heterostructures. Adv. Mater. 2016, 28, 3742–3748. [Google Scholar] [CrossRef] [PubMed]
  169. Shih, C.J.; Wang, Q.H.; Son, Y.; Jin, Z.; Blankschtein, D.; Strano, M.S. Tuning On-Off Current Ratio and Field-Effect Mobility in a MoS2-Graphene Heterostructure via Schottky Barrier Modulation. ACS Nano 2014, 8, 5790–5798. [Google Scholar] [CrossRef] [PubMed]
  170. Kwak, J.Y.; Hwang, J.; Calderon, B.; Alsalman, H.; Munoz, N.; Schutter, B.; Spencer, M.G. Electrical characteristics of multilayer MoS2 FET's with MoS2/graphene heterojunction contacts. Nano Lett. 2014, 14, 4511–4516. [Google Scholar] [CrossRef] [PubMed]
  171. Moriya, R.; Yamaguchi, T.; Inoue, Y.; Morikawa, S.; Sata, Y.; Masubuchi, S.; Machida, T. Large current modulation in exfoliated-graphene/MoS2/metal vertical heterostructures. Appl. Phys. Lett. 2014, 105, 083119. [Google Scholar] [CrossRef]
  172. Lin, Y.F.; Li, W.; Li, S.L.; Xu, Y.; Aparecido-Ferreira, A.; Komatsu, K.; Sun, H.; Nakaharai, S.; Tsukagoshi, K. Barrier inhomogeneities at vertically stacked graphene-based heterostructures. Nanoscale 2014, 6, 795–799. [Google Scholar] [CrossRef] [PubMed]
  173. Qiu, D.; Kim, E.K. Electrically Tunable and Negative Schottky Barriers in Multi-layered Graphene/MoS2 Heterostructured Transistors. Sci. Rep. 2015, 5, 13743. [Google Scholar] [CrossRef] [PubMed]
  174. Cheng, H.C.; Wang, G.; Li, D.; He, Q.; Yin, A.; Liu, Y.; Wu, H.; Ding, M.; Huang, Y.; Duan, X. van der Waals Heterojunction Devices Based on Organohalide Perovskites and Two-Dimensional Materials. Nano Lett. 2016, 16, 367–373. [Google Scholar] [CrossRef] [PubMed]
  175. Shim, J.; Kim, H.S.; Shim, Y.S.; Kang, D.H.; Park, H.Y.; Lee, J.; Jeon, J.; Jung, S.J.; Song, Y.J.; Jung, W.S.; et al. Extremely Large Gate Modulation in Vertical Graphene/WSe2 Heterojunction Barristor Based on a Novel Transport Mechanism. Adv. Mater. 2016, 28, 5293–5299. [Google Scholar] [CrossRef] [PubMed]
  176. Bernardi, M.; Palummo, M.; Grossman, J.C. Extraordinary sunlight absorption and one nanometer thick photovoltaics using two-dimensional monolayer materials. Nano Lett. 2013, 13, 3664–3670. [Google Scholar] [CrossRef] [PubMed]
  177. Britnell, L.; Ribeiro, R.M.; Eckmann, A.; Jalil, R.; Belle, B.D.; Mishchenko, A.; Kim, Y.J.; Gorbachev, R.V.; Georgiou, T.; Morozov, S.V.; et al. Strong light-matter interactions in heterostructures of atomically thin films. Science 2013, 340, 1311–1314. [Google Scholar] [CrossRef] [PubMed]
  178. Eda, G.; Maier, S.A. Two-dimensional crystals: Managing light for optoelectronics. ACS Nano 2013, 7, 5660–5665. [Google Scholar] [CrossRef] [PubMed]
  179. Yu, W.J.; Liu, Y.; Zhou, H.; Yin, A.; Li, Z.; Huang, Y.; Duan, X. Highly efficient gate-tunable photocurrent generation in vertical heterostructures of layered materials. Nat. Nanotechnol. 2013, 8, 952–958. [Google Scholar] [CrossRef] [PubMed]
  180. Massicotte, M.; Schmidt, P.; Vialla, F.; Schadler, K.G.; Reserbat-Plantey, A.; Watanabe, K.; Taniguchi, T.; Tielrooij, K.J.; Koppens, F.H. Picosecond photoresponse in van der Waals heterostructures. Nat. Nanotechnol. 2016, 11, 42–46. [Google Scholar] [CrossRef] [PubMed]
  181. Long, M.; Liu, E.; Wang, P.; Gao, A.; Xia, H.; Luo, W.; Wang, B.; Zeng, J.; Fu, Y.; Xu, K.; et al. Broadband Photovoltaic Detectors Based on an Atomically Thin Heterostructure. Nano Lett. 2016, 16, 2254–2259. [Google Scholar] [CrossRef] [PubMed]
  182. Furchi, M.M.; Pospischil, A.; Libisch, F.; Burgdorfer, J.; Mueller, T. Photovoltaic effect in an electrically tunable van der Waals heterojunction. Nano Lett. 2014, 14, 4785–4791. [Google Scholar] [CrossRef] [PubMed]
  183. Lee, C.H.; Lee, G.H.; van der Zande, A.M.; Chen, W.; Li, Y.; Han, M.; Cui, X.; Arefe, G.; Nuckolls, C.; Heinz, T.F.; et al. Atomically thin p-n junctions with van der Waals heterointerfaces. Nat. Nanotechnol. 2014, 9, 676–681. [Google Scholar] [CrossRef] [PubMed]
  184. Cheng, R.; Li, D.; Zhou, H.; Wang, C.; Yin, A.; Jiang, S.; Liu, Y.; Chen, Y.; Huang, Y.; Duan, X. Electroluminescence and photocurrent generation from atomically sharp WSe2/MoS2 heterojunction p-n diodes. Nano Lett. 2014, 14, 5590–5597. [Google Scholar] [CrossRef] [PubMed]
  185. Deng, Y.X.; Luo, Z.; Conrad, N.J.; Liu, H.; Gong, Y.J.; Najmaei, S.; Ajayan, P.M.; Lou, J.; Xu, X.F.; Ye, P.D. Black Phosphorus-Monolayer MoS2 van der Waals Heterojunction p–n Diode. ACS Nano 2014, 8, 8292–8299. [Google Scholar] [CrossRef] [PubMed]
  186. Wang, Q.H.; Kalantar-Zadeh, K.; Kis, A.; Coleman, J.N.; Strano, M.S. Electronics and optoelectronics of two-dimensional transition metal dichalcogenides. Nat. Nanotechnol. 2012, 7, 699–712. [Google Scholar] [CrossRef] [PubMed]
  187. Bertolazzi, S.; Brivio, J.; Kis, A. Stretching and Breaking of Ultrathin MoS2. Acs Nano 2011, 5, 9703–9709. [Google Scholar] [CrossRef] [PubMed]
  188. Jariwala, D.; Sangwan, V.K.; Lauhon, L.J.; Marks, T.J.; Hersam, M.C. Emerging device applications for semiconducting two-dimensional transition metal dichalcogenides. ACS Nano 2014, 8, 1102–1120. [Google Scholar] [CrossRef] [PubMed]
  189. Lopez-Sanchez, O.; Alarcon Llado, E.; Koman, V.; Fontcuberta i Morral, A.; Radenovic, A.; Kis, A. Light generation and harvesting in a van der Waals heterostructure. ACS Nano 2014, 8, 3042–3048. [Google Scholar] [CrossRef] [PubMed]
  190. Esmaeili-Rad, M.R.; Salahuddin, S. High performance molybdenum disulfide amorphous silicon heterojunction photodetector. Sci. Rep. 2013, 3, 2345. [Google Scholar] [CrossRef] [PubMed]
  191. Gourmelon, E.; Lignier, O.; Hadouda, H.; Couturier, G.; Bernwde, J.C.; Tedd, J.; Pouzet, J.; Salardenne, J. MS2 (M = W, Mo) Photosensitive thin film for solar cells. Sol. Energy Mater. Sol. Cells 1997, 46, 115–121. [Google Scholar] [CrossRef]
  192. Thomalla, M.; Tributsch, H. Photosensitization of nanostructured TiO2 with WS2 quantum sheets. J. Phys. Chem. B 2006, 110, 12167–12171. [Google Scholar] [CrossRef] [PubMed]
  193. Shanmugam, M.; Bansal, T.; Durcan, C.A.; Yu, B. Molybdenum disulphide/titanium dioxide nanocomposite-poly 3-hexylthiophene bulk heterojunction solar cell. Appl. Phys. Lett. 2012, 100, 153901. [Google Scholar] [CrossRef]
  194. Shanmugam, M.; Durcan, C.A.; Yu, B. Layered semiconductor molybdenum disulfide nanomembrane based Schottky-barrier solar cells. Nanoscale 2012, 4, 7399–7405. [Google Scholar] [CrossRef] [PubMed]
  195. Wi, S.; Kim, H.; Chen, M.; Nam, H.; Guo, L.J.; Meyhofer, E.; Liang, X. Enhancement of photovoltaic response in multilayer MoS2 induced by plasma doping. ACS Nano 2014, 8, 5270–5281. [Google Scholar] [CrossRef] [PubMed]
  196. Polman, A.; Atwater, H.A. Photonic design principles for ultrahigh-efficiency photovoltaics. Nat. Mater. 2012, 11, 174–177. [Google Scholar] [CrossRef] [PubMed]
  197. Lin, J.D.; Li, H.; Zhang, H.; Chen, W. Plasmonic enhancement of photocurrent in MoS2 field-effect-transistor. Appl. Phys. Lett. 2013, 102, 203109. [Google Scholar] [CrossRef]
  198. Lee, H.S.; Min, S.W.; Chang, Y.G.; Park, M.K.; Nam, T.; Kim, H.; Kim, J.H.; Ryu, S.; Im, S. MoS2 nanosheet phototransistors with thickness-modulated optical energy gap. Nano Lett. 2012, 12, 3695–3700. [Google Scholar] [CrossRef] [PubMed]
  199. Feng, J.; Qian, X.F.; Huang, C.W.; Li, J. Strain-engineered artificial atom as a broad-spectrum solar energy funnel. Nat. Photonics 2012, 6, 865–871. [Google Scholar] [CrossRef]
  200. Carladous, A.; Coratger, R.; Ajustron, F.; Seine, G.; Péchou, R.; Beauvillain, J. Light emission from spectral analysis of Au/MoS2 nanocontacts stimulated by scanning tunneling microscopy. Phys. Rev. B 2002, 66, 045401. [Google Scholar] [CrossRef]
  201. Kirmayer, S.; Aharon, E.; Dovgolevsky, E.; Kalina, M.; Frey, G.L. Self-assembled lamellar MoS2, SnS2 and SiO2 semiconducting polymer nanocomposites. Philos. Trans. A Math. Phys. Eng. Sci. 2007, 365, 1489–1508. [Google Scholar] [CrossRef] [PubMed]
  202. Sundaram, R.S.; Engel, M.; Lombardo, A.; Krupke, R.; Ferrari, A.C.; Avouris, P.; Steiner, M. Electroluminescence in single layer MoS2. Nano Lett. 2013, 13, 1416–1421. [Google Scholar] [CrossRef] [PubMed]
  203. Zhang, Y.J.; Oka, T.; Suzuki, R.; Ye, J.T.; Iwasa, Y. Electrically Switchable Chiral Light-Emitting Transistor. Science 2014, 344, 725–728. [Google Scholar] [CrossRef] [PubMed]
  204. Withers, F.; Del Pozo-Zamudio, O.; Mishchenko, A.; Rooney, A.P.; Gholinia, A.; Watanabe, K.; Taniguchi, T.; Haigh, S.J.; Geim, A.K.; Tartakovskii, A.I.; et al. Light-emitting diodes by band-structure engineering in van der Waals heterostructures. Nat. Mater. 2015, 14, 301–306. [Google Scholar] [CrossRef] [PubMed]
  205. Withers, F.; Del Pozo-Zamudio, O.; Schwarz, S.; Dufferwiel, S.; Walker, P.M.; Godde, T.; Rooney, A.P.; Gholinia, A.; Woods, C.R.; Blake, P.; et al. WSe(2) Light-Emitting Tunneling Transistors with Enhanced Brightness at Room Temperature. Nano Lett. 2015, 15, 8223–8228. [Google Scholar] [CrossRef] [PubMed]
  206. Muccini, M.; Toffanin, S. Organic Light-Emitting Transistors, in Organic Light-Emitting Transistors; John Wiley & Sons, Inc.: Hoboken, NJ, USA, 2016; pp. 45–85. [Google Scholar]
  207. Clark, G.; Schaibley, J.R.; Ross, J.; Taniguchi, T.; Watanabe, K.; Hendrickson, J.R.; Mou, S.; Yao, W.; Xu, X. Single Defect Light-Emitting Diode in a van der Waals Heterostructure. Nano Lett. 2016, 16, 3944–3948. [Google Scholar] [CrossRef] [PubMed]
  208. Wang, H.; Zhang, C.; Chan, W.; Tiwari, S.; Rana, F. Ultrafast response of monolayer molybdenum disulfide photodetectors. Nat. Commun. 2015, 6, 8831. [Google Scholar] [CrossRef] [PubMed]
  209. Lopez-Sanchez, O.; Lembke, D.; Kayci, M.; Radenovic, A.; Kis, A. Ultrasensitive photodetectors based on monolayer MoS2. Nat. Nanotechnol. 2013, 8, 497–501. [Google Scholar] [CrossRef] [PubMed]
  210. Zhang, W.; Huang, J.K.; Chen, C.H.; Chang, Y.H.; Cheng, Y.J.; Li, L.J. High-gain phototransistors based on a CVD MoS(2) monolayer. Adv. Mater. 2013, 25, 3456–3461. [Google Scholar] [CrossRef] [PubMed]
  211. Perea-López, N.; Elías, A.L.; Berkdemir, A.; Castro-Beltran, A.; Gutiérrez, H.R.; Feng, S.; Lv, R.; Hayashi, T.; López-Urías, F.; Ghosh, S.; et al. Photosensor Device Based on Few-Layered WS2Films. Adv. Funct. Mater. 2013, 23, 5511–5517. [Google Scholar] [CrossRef]
  212. Zhang, W.; Chiu, M.H.; Chen, C.H.; Chen, W.; Li, L.J.; Wee, A.T. Role of metal contacts in high-performance phototransistors based on WSe2 monolayers. ACS Nano 2014, 8, 8653–8661. [Google Scholar] [CrossRef] [PubMed]
  213. Tsai, D.S.; Liu, K.K.; Lien, D.H.; Tsai, M.L.; Kang, C.F.; Lin, C.A.; Li, L.J.; He, J.H. Few-Layer MoS2 with High Broadband Photogain and Fast Optical Switching for Use in Harsh Environments. ACS Nano 2013, 7, 3905–3911. [Google Scholar] [CrossRef] [PubMed]
  214. Dung-Sheng, T.; Lien, D.H.; Tsai, M.L.; Su, S.H.; Chen, K.M.; Ke, J.J.; Yu, Y.C.; Li, L.J.; He, J.H. Trilayered MoS2 Metal -Semiconductor-Metal Photodetectors: Photogain and Radiation Resistance. IEEE J. Sel. Top. Quantum. Electron. 2014, 20, 30–35. [Google Scholar] [CrossRef]
  215. Yamaguchi, H.; Blancon, J.C.; Kappera, R.; Lei, S.; Najmaei, S.; Mangum, B.D.; Gupta, G.; Ajayan, P.M.; Lou, J.; Chhowalla, M.; et al. Spatially resolved photoexcited charge-carrier dynamics in phase-engineered monolayer MoS2. ACS Nano 2015, 9, 840–849. [Google Scholar] [CrossRef] [PubMed]
  216. Kang, D.-H.; Kim, M.S.; Shim, J.; Jeon, J.; Park, H.Y.; Jung, W.S.; Yu, H.Y.; Pang, C.H.; Lee, S.; Park, J.H. High-Performance Transition Metal Dichalcogenide Photodetectors Enhanced by Self-Assembled Monolayer Doping. Adv. Funct. Mater. 2015, 25, 4219–4227. [Google Scholar] [CrossRef]
  217. Gan, X.T.; Shiue, R.J.; Gao, Y.D.; Meric, I.; Heinz, T.F.; Shepard, K.; Hone, J.; Assefa, S.; Englund, D. Chip-integrated ultrafast graphene photodetector with high responsivity. Nat. Photonics 2013, 7, 883–887. [Google Scholar] [CrossRef]
  218. Mittendorff, M.; Winnerl, S.; Kamann, J.; Eroms, J.; Weiss, D.; Schneider, H.; Helm, M. Ultrafast graphene-based broadband THz detector. Appl. Phys. Lett. 2013, 103, 021113. [Google Scholar] [CrossRef]
  219. Xu, X.; Gabor, N.M.; Alden, J.S.; van der Zande, A.M.; McEuen, P.L. Photo-thermoelectric effect at a graphene interface junction. Nano Lett. 2010, 10, 562–566. [Google Scholar] [CrossRef] [PubMed]
  220. Zhang, Y.; Li, H.; Wang, L.; Wang, H.; Xie, X.; Zhang, S.L.; Liu, R.; Qiu, Z.J. Photothermoelectric and photovoltaic effects both present in MoS2. Sci. Rep. 2015, 5, 7938. [Google Scholar] [CrossRef] [PubMed]
  221. Buscema, M.; Barkelid, M.; Zwiller, V.; van der Zant, H.S.; Steele, G.A.; Castellanos-Gomez, A. Large and tunable photothermoelectric effect in single-layer MoS2. Nano Lett. 2013, 13, 358–363. [Google Scholar] [CrossRef] [PubMed]
  222. Konabe, S.; Yamamoto, T. Valley photothermoelectric effects in transition-metal dichalcogenides. Phys. Rev. B 2014, 90, 075430. [Google Scholar] [CrossRef]
  223. Choi, W.; Cho, M.Y.; Konar, A.; Lee, J.H.; Cha, G.B.; Hong, S.C.; Kim, S.; Kim, J.; Jena, D.; Joo, J.; et al. High-detectivity multilayer MoS(2) phototransistors with spectral response from ultraviolet to infrared. Adv. Mater. 2012, 24, 5832–5836. [Google Scholar] [CrossRef] [PubMed]
  224. Kim, H.C.; Kim, H.; Lee, J.U.; Lee, H.B.; Choi, D.H.; Lee, J.H.; Lee, W.H.; Jhang, S.H.; Park, B.H.; Cheong, H.; et al. Engineering Optical and Electronic Properties of WS2 by Varying the Number of Layers. ACS Nano 2015, 9, 6854–6860. [Google Scholar] [CrossRef] [PubMed]
  225. Jeong, H.; Oh, H.M.; Bang, S.; Jeong, H.J.; An, S.J.; Han, G.H.; Kim, H.; Yun, S.J.; Kim, K.K.; Park, J.C.; et al. Metal-Insulator-Semiconductor Diode Consisting of Two-Dimensional Nanomaterials. Nano Lett. 2016, 16, 1858–1862. [Google Scholar] [CrossRef] [PubMed]
  226. Yu, S.H.; Lee, Y.; Jang, S.K.; Kang, J.; Jeon, J.; Lee, C.; Lee, J.Y.; Kim, H.; Hwang, E.; Lee, S.; et al. Dye-sensitized MoS2 photodetector with enhanced spectral photoresponse. ACS Nano 2014, 8, 8285–8291. [Google Scholar] [CrossRef] [PubMed]
  227. Kufer, D.; Nikitskiy, I.; Lasanta, T.; Navickaite, G.; Koppens, F.H.; Konstantatos, G. Hybrid 2D-0D MoS2 -PbS quantum dot photodetectors. Adv. Mater. 2015, 27, 176–180. [Google Scholar] [CrossRef] [PubMed]
  228. Kang, D.H.; Pae, S.R.; Shim, J.; Yoo, G.; Jeon, J.; Leem, J.W.; Yu, J.S.; Lee, S.; Shin, B.; Park, J.H. An Ultrahigh-Performance Photodetector based on a Perovskite-Transition-Metal-Dichalcogenide Hybrid Structure. Adv. Mater. 2016, 28, 7799–7806. [Google Scholar] [CrossRef] [PubMed]
  229. Splendiani, A.; Sun, L.; Zhang, Y.; Li, T.; Kim, J.; Chim, C.Y.; Galli, G.; Wang, F. Emerging photoluminescence in monolayer MoS2. Nano Lett. 2010, 10, 1271–1275. [Google Scholar] [CrossRef] [PubMed]
  230. Eda, G.; Yamaguchi, H.; Voiry, D.; Fujita, T.; Chen, M.; Chhowalla, M. Photoluminescence from chemically exfoliated MoS2. Nano Lett. 2011, 11, 5111–5116. [Google Scholar] [CrossRef] [PubMed]
  231. Gutierrez, H.R.; Perea-Lopez, N.; Elias, A.L.; Berkdemir, A.; Wang, B.; Lv, R.; Lopez-Urias, F.; Crespi, V.H.; Terrones, H.; Terrones, M. Extraordinary room-temperature photoluminescence in triangular WS2 monolayers. Nano Lett. 2013, 13, 3447–3454. [Google Scholar] [CrossRef] [PubMed]
  232. Tonndorf, P.; Schmidt, R.; Bottger, P.; Zhang, X.; Borner, J.; Liebig, A.; Albrecht, M.; Kloc, C.; Gordan, O.; Zahn, D.R.T.; et al. Photoluminescence emission and Raman response of monolayer MoS2, MoSe2, and WSe2. Opt. Express 2013, 21, 4908–4916. [Google Scholar] [CrossRef] [PubMed]
  233. Mak, K.F.; Lee, C.; Hone, J.; Shan, J.; Heinz, T.F. Atomically Thin MoS2: A New Direct-Gap Semiconductor. Phys. Rev. Lett. 2010, 105, 136805. [Google Scholar] [CrossRef] [PubMed]
  234. Salehzadeh, O.; Tran, N.H.; Liu, X.; Shih, I.; Mi, Z. Exciton kinetics, quantum efficiency, and efficiency droop of monolayer MoS(2) light-emitting devices. Nano Lett. 2014, 14, 4125–4130. [Google Scholar] [CrossRef] [PubMed]
  235. Salehzadeh, O.; Djavid, M.; Tran, N.H.; Shih, I.; Mi, Z. Optically Pumped Two-Dimensional MoS2 Lasers Operating at Room-Temperature. Nano Lett. 2015, 15, 5302–5306. [Google Scholar] [CrossRef] [PubMed]
  236. Gan, X.; Gao, Y.; Fai Mak, K.; Yao, X.; Shiue, R.J.; van der Zande, A.; Trusheim, M.E.; Hatami, F.; Heinz, T.F.; Hone, J.; et al. Controlling the spontaneous emission rate of monolayer MoS2 in a photonic crystal nanocavity. Appl. Phys. Lett. 2013, 103, 181119. [Google Scholar] [CrossRef] [PubMed]
  237. Wu, S.F.; Buckley, S.; Jones, A.M.; Ross, J.S.; Ghimire, N.J.; Yan, J.Q.; Mandrus, D.G.; Yao, W.; Hatami, F.; Vuckovic, J.; et al. Control of two-dimensional excitonic light emission via photonic crystal. 2D Mater. 2014, 1, 011001. [Google Scholar] [CrossRef]
  238. Schwarz, S.; Dufferwiel, S.; Walker, P.M.; Withers, F.; Trichet, A.A.; Sich, M.; Li, F.; Chekhovich, E.A.; Borisenko, D.N.; Kolesnikov, N.N.; et al. Two-dimensional metal-chalcogenide films in tunable optical microcavities. Nano Lett. 2014, 14, 7003–7008. [Google Scholar] [CrossRef] [PubMed]
  239. Liu, X.; Galfsky, T.; Sun, Z.; Xia, F.; Lin, E.-C.; Lee, Y.-H.; Kéna-Cohen, S.; Menon, V.M. Strong light–matter coupling in two-dimensional atomic crystals. Nat. Photonics 2014, 9, 30–34. [Google Scholar] [CrossRef]
  240. Wu, S.; Buckley, S.; Schaibley, J.R.; Feng, L.; Yan, J.; Mandrus, D.G.; Hatami, F.; Yao, W.; Vuckovic, J.; Majumdar, A.; et al. Monolayer semiconductor nanocavity lasers with ultralow thresholds. Nature 2015, 520, 69–72. [Google Scholar] [CrossRef] [PubMed]
  241. Mak, K.F.; He, K.; Lee, C.; Lee, G.H.; Hone, J.; Heinz, T.F.; Shan, J. Tightly bound trions in monolayer MoS2. Nat. Mater. 2013, 12, 207–211. [Google Scholar] [CrossRef] [PubMed]
  242. Ye, Z.; Cao, T.; O’Brien, K.; Zhu, H.; Yin, X.; Wang, Y.; Louie, S.G.; Zhang, X. Probing excitonic dark states in single-layer tungsten disulphide. Nature 2014, 513, 214–218. [Google Scholar] [CrossRef] [PubMed]
  243. He, K.; Kumar, N.; Zhao, L.; Wang, Z.; Mak, K.F.; Zhao, H.; Shan, J. Tightly bound excitons in monolayer WSe(2). Phys. Rev. Lett. 2014, 113, 026803. [Google Scholar] [CrossRef] [PubMed]
  244. Ye, Y.; Wong, Z.J.; Lu, X.F.; Ni, X.J.; Zhu, H.Y.; Chen, X.H.; Wang, Y.; Zhang, X. Monolayer excitonic laser. Nat. Photonics 2015, 9, 733–737. [Google Scholar] [CrossRef]
  245. Song, H.S.; Li, S.L.; Gao, L.; Xu, Y.; Ueno, K.; Tang, J.; Cheng, Y.B.; Tsukagoshi, K. High-performance top-gated monolayer SnS2 field-effect transistors and their integrated logic circuits. Nanoscale 2013, 5, 9666–9670. [Google Scholar] [CrossRef] [PubMed]
  246. Tosun, M.; Chuang, S.; Fang, H.; Sachid, A.B.; Hettick, M.; Lin, Y.; Zeng, Y.; Javey, A. High-gain inverters based on WSe2 complementary field-effect transistors. ACS Nano 2014, 8, 4948–4953. [Google Scholar] [CrossRef] [PubMed]
  247. Cho, A.J.; Park, K.C.; Kwon, J.Y. A high-performance complementary inverter based on transition metal dichalcogenide field-effect transistors. Nanoscale Res. Lett. 2015, 10, 115. [Google Scholar] [CrossRef] [PubMed]
  248. Jeon, P.J.; Kim, J.S.; Lim, J.Y.; Cho, Y.; Pezeshki, A.; Lee, H.S.; Yu, S.; Min, S.W.; Im, S. Low Power Consumption Complementary Inverters with n-MoS2 and p-WSe2 Dichalcogenide Nanosheets on Glass for Logic and Light-Emitting Diode Circuits. ACS Appl. Mater. Interfaces 2015, 7, 22333–22340. [Google Scholar] [CrossRef] [PubMed]
  249. Wang, H.; Yu, L.; Lee, Y.H.; Shi, Y.; Hsu, A.; Chin, M.L.; Li, L.J.; Dubey, M.; Kong, J.; Palacios, T. Integrated circuits based on bilayer MoS(2) transistors. Nano Lett. 2012, 12, 4674–4680. [Google Scholar] [CrossRef] [PubMed]
  250. Lipatov, A.; Sharma, P.; Gruverman, A.; Sinitskii, A. Optoelectrical Molybdenum Disulfide (MoS2)-Ferroelectric Memories. ACS Nano 2015, 9, 8089–8098. [Google Scholar] [CrossRef] [PubMed]
  251. Zhang, E.; Wang, W.; Zhang, C.; Jin, Y.; Zhu, G.; Sun, Q.; Zhang, D.W.; Zhou, P.; Xiu, F. Tunable charge-trap memory based on few-layer MoS2. ACS Nano 2015, 9, 612–619. [Google Scholar] [CrossRef] [PubMed]
  252. Bertolazzi, S.; Krasnozhon, D.; Kis, A. Nonvolatile memory cells based on MoS2/graphene heterostructures. ACS Nano 2013, 7, 3246–3252. [Google Scholar] [CrossRef] [PubMed]
  253. Choi, M.S.; Lee, G.H.; Yu, Y.J.; Lee, D.Y.; Lee, S.H.; Kim, P.; Hone, J.; Yoo, W.J. Controlled charge trapping by molybdenum disulphide and graphene in ultrathin heterostructured memory devices. Nat. Commun. 2013, 4, 1624. [Google Scholar] [CrossRef] [PubMed]
  254. Vu, Q.A.; Shin, Y.S.; Kim, Y.R.; Nguyen, V.L.; Kang, W.T.; Kim, H.; Luong, D.H.; Lee, I.M.; Lee, K.; Ko, D.S.; et al. Two-terminal floating-gate memory with van der Waals heterostructures for ultrahigh on/off ratio. Nat. Commun. 2016, 7, 12725. [Google Scholar] [CrossRef] [PubMed]
  255. Roy, K.; Padmanabhan, M.; Goswami, S.; Sai, T.P.; Ramalingam, G.; Raghavan, S.; Ghosh, A. Graphene-MoS2 hybrid structures for multifunctional photoresponsive memory devices. Nat. Nanotechnol. 2013, 8, 826–830. [Google Scholar] [CrossRef] [PubMed]
  256. Lee, Y.T.; Lee, J.; Ju, H.; Lim, J.A.; Yi, Y.; Choi, W.K.; Hwang, D.K.; Im, S. Nonvolatile Charge Injection Memory Based on Black Phosphorous 2D Nanosheets for Charge Trapping and Active Channel Layers. Adv. Funct. Mater. 2016, 26, 5701–5707. [Google Scholar] [CrossRef]
  257. Liu, H.L.; Shen, C.C.; Su, S.H.; Hsu, C.L.; Li, M.Y.; Li, L.J. Optical properties of monolayer transition metal dichalcogenides probed by spectroscopic ellipsometry. Appl. Phys. Lett. 2014, 105, 201905. [Google Scholar] [CrossRef]
Figure 1. (a) A MoS2 thin film transistor and its transport properties. (b) Drain current (ID) versus back-gate bias (VGS) with different drain bias (VDS). (c) Drain current (ID) versus drain bias (VDS) showing current saturation, including a long-channel model (red lines) showing excellent agreement with measured device behavior. Reproduced with permission from [46], Nature Publishing Group, 2012.
Figure 1. (a) A MoS2 thin film transistor and its transport properties. (b) Drain current (ID) versus back-gate bias (VGS) with different drain bias (VDS). (c) Drain current (ID) versus drain bias (VDS) showing current saturation, including a long-channel model (red lines) showing excellent agreement with measured device behavior. Reproduced with permission from [46], Nature Publishing Group, 2012.
Electronics 06 00043 g001
Figure 2. Charge transport properties of the multilayer MoS2 channel. Measured temperature-dependent mobility of MoS2 field effect transistors (FETs). Reproduced with permission from [46], Nature Publishing Group, 2012.
Figure 2. Charge transport properties of the multilayer MoS2 channel. Measured temperature-dependent mobility of MoS2 field effect transistors (FETs). Reproduced with permission from [46], Nature Publishing Group, 2012.
Electronics 06 00043 g002
Figure 3. Current distribution among the individual layers of a 13-layer thick MoS2 FET at different gate bias conditions derived using the resistor network model. The illustrations show the location of the “HOT-SPOT” and the associated current spread schematically corresponding to two different gate bias conditions. The red and blue colors representing the highest and lowest current, respectively. Reproduced with permission from [89], American Chemical Society, 2013.
Figure 3. Current distribution among the individual layers of a 13-layer thick MoS2 FET at different gate bias conditions derived using the resistor network model. The illustrations show the location of the “HOT-SPOT” and the associated current spread schematically corresponding to two different gate bias conditions. The red and blue colors representing the highest and lowest current, respectively. Reproduced with permission from [89], American Chemical Society, 2013.
Electronics 06 00043 g003
Figure 4. The temperature dependence of the mobility at carrier density n = 1011 cm−2 calculated with the full collision integral. For comparison, the mobility in the presence of only acoustic deformation potential scattering with the temperature dependence is shown by blue line. The (gray) shaded area shows the variation in the mobility associated with a 10% uncertainty in the calculated deformation potentials. Reproduced with permission from [120], American Physical Society, 2012.
Figure 4. The temperature dependence of the mobility at carrier density n = 1011 cm−2 calculated with the full collision integral. For comparison, the mobility in the presence of only acoustic deformation potential scattering with the temperature dependence is shown by blue line. The (gray) shaded area shows the variation in the mobility associated with a 10% uncertainty in the calculated deformation potentials. Reproduced with permission from [120], American Physical Society, 2012.
Electronics 06 00043 g004
Figure 5. (a) Three-dimensional schematic view of the monolayer MoS2 transistors. Top gate width and length are 4 um and 500 nm, respectively. (b) Ids–Vtg curve recorded for Vds ranging from 10 to 500 mV. Measurements were performed at room temperature with the back-gate grounded. The device can be completely turned off by changing the top gate bias from −2 to −4 V. For Vds = 10 mV, the Ion/Ioff is > 1 × 106. For Vds = 500 mV, the Ion/Ioff is > 1 × 108 while the subthreshold swing 74 mV/dec. Inset: Ids–Vtg for values of Vbg = −10, −5, 0, 5, and 10 V. Reproduced with permission from [82], Nature Publishing Group, 2011.
Figure 5. (a) Three-dimensional schematic view of the monolayer MoS2 transistors. Top gate width and length are 4 um and 500 nm, respectively. (b) Ids–Vtg curve recorded for Vds ranging from 10 to 500 mV. Measurements were performed at room temperature with the back-gate grounded. The device can be completely turned off by changing the top gate bias from −2 to −4 V. For Vds = 10 mV, the Ion/Ioff is > 1 × 106. For Vds = 500 mV, the Ion/Ioff is > 1 × 108 while the subthreshold swing 74 mV/dec. Inset: Ids–Vtg for values of Vbg = −10, −5, 0, 5, and 10 V. Reproduced with permission from [82], Nature Publishing Group, 2011.
Electronics 06 00043 g005
Figure 6. Performance of a high-transconductance MoS2 FET. (a) Ids–Vds characteristics measured for different top-gate voltages Vtg from 0 to 6 V. Inset: Three-dimensional representation of the monolayer MoS2 FET (b) Transconductance derived from Ids–Vds characteristics at Vds = 0.5 V, 1 V, and 4 V. Reproduced with permission from [103], American Chemical Society, 2012.
Figure 6. Performance of a high-transconductance MoS2 FET. (a) Ids–Vds characteristics measured for different top-gate voltages Vtg from 0 to 6 V. Inset: Three-dimensional representation of the monolayer MoS2 FET (b) Transconductance derived from Ids–Vds characteristics at Vds = 0.5 V, 1 V, and 4 V. Reproduced with permission from [103], American Chemical Society, 2012.
Electronics 06 00043 g006
Figure 7. (a) Hysteresis in the monolayer MoS2 transistor with controlled humidity under the sweeping rate of gate voltage at 0.5 V/s and Vds = 0.5 V. (b) Hysteresis with different illumination conditions in a monolayer MoS2 transistor. Blue dots are under global white illumination (0.7 mW/cm2), and red dots are in the dark. The intersections between the dashed lines, and y = 0 show the threshold voltages. Reproduced with permission from [99], American Chemical Society, 2012.
Figure 7. (a) Hysteresis in the monolayer MoS2 transistor with controlled humidity under the sweeping rate of gate voltage at 0.5 V/s and Vds = 0.5 V. (b) Hysteresis with different illumination conditions in a monolayer MoS2 transistor. Blue dots are under global white illumination (0.7 mW/cm2), and red dots are in the dark. The intersections between the dashed lines, and y = 0 show the threshold voltages. Reproduced with permission from [99], American Chemical Society, 2012.
Electronics 06 00043 g007
Figure 8. Transfer curve of bulk (1000 μm × 500 μm × 10 μm) and thin-flake (20 μm × 20 μm × 15 nm) MoS2 EDLTs. (a) Change in the channel current (IDS) as a function of gate voltage (VG). (b) Change of sheet conductivity (σ2D) as a function of gate voltage (VG). Channel voltage (VDS) is 0.2 V for both devices. Reproduced with permission from [150], American Chemical Society, 2012.
Figure 8. Transfer curve of bulk (1000 μm × 500 μm × 10 μm) and thin-flake (20 μm × 20 μm × 15 nm) MoS2 EDLTs. (a) Change in the channel current (IDS) as a function of gate voltage (VG). (b) Change of sheet conductivity (σ2D) as a function of gate voltage (VG). Channel voltage (VDS) is 0.2 V for both devices. Reproduced with permission from [150], American Chemical Society, 2012.
Electronics 06 00043 g008
Figure 9. Schematic illustration of the structure and working principle of a hexagonal boron nitride (h-BN)/WSe2 FET with ionic-liquid-gated graphene contacts. Reproduced with permission from [69], American Chemical Society, 2014.
Figure 9. Schematic illustration of the structure and working principle of a hexagonal boron nitride (h-BN)/WSe2 FET with ionic-liquid-gated graphene contacts. Reproduced with permission from [69], American Chemical Society, 2014.
Electronics 06 00043 g009
Figure 10. (a) Schematic of the MoTe2 ambipolar transistor structure. The gap between the two top gates is 100 nm. (b) Experimental results for on/off operation in the p-FET mode (VtgS = −5 V). The inset is a logarithmic plot. (c) Experimental results of an on/off operation in the n-FET mode (VtgS = 5 V), with a logarithmic plot provided in the inset. Reproduced with permission from [160], American Chemical Society, 2015.
Figure 10. (a) Schematic of the MoTe2 ambipolar transistor structure. The gap between the two top gates is 100 nm. (b) Experimental results for on/off operation in the p-FET mode (VtgS = −5 V). The inset is a logarithmic plot. (c) Experimental results of an on/off operation in the n-FET mode (VtgS = 5 V), with a logarithmic plot provided in the inset. Reproduced with permission from [160], American Chemical Society, 2015.
Electronics 06 00043 g010
Figure 11. Bandgap vs. field-effect mobility of some important semiconductors used in optoelectronic devices. The dashed line is the band gap at ~1.34 eV while efficiency reach the Shockley-Quessier limit (~33.7%). Reproduced with permission from [188], American Chemical Society, 2014.
Figure 11. Bandgap vs. field-effect mobility of some important semiconductors used in optoelectronic devices. The dashed line is the band gap at ~1.34 eV while efficiency reach the Shockley-Quessier limit (~33.7%). Reproduced with permission from [188], American Chemical Society, 2014.
Electronics 06 00043 g011
Figure 12. Bandgap absorbance of three TMD monolayers and graphene, overlapped to the incident AM1.5G solar flux. Reproduced with permission from [176], American Chemical Society, 2013.
Figure 12. Bandgap absorbance of three TMD monolayers and graphene, overlapped to the incident AM1.5G solar flux. Reproduced with permission from [176], American Chemical Society, 2013.
Electronics 06 00043 g012
Figure 13. I–V characteristics of monolayer WSe2 photodiode under a halogen lamp with 1400 W/m2. The biasing conditions are p–n (solid green line; VG1 = −40 V, VG2 = 40 V), n–p (solid blue line; VG1 = 40 V, VG2 = −40 V), and p–p (dash blue line; VG1 = VG2 = 40 V). The red dashed rectangle in the main panel shows the maximum power conversion efficiency area (Pel). Top inset: Schematic of WSe2 monolayer device with split gate electrodes. Lower inset: electrical power (Pel) versus voltage under incident illumination. Reproduced with permission from [157], Nature Publishing Group, 2014.
Figure 13. I–V characteristics of monolayer WSe2 photodiode under a halogen lamp with 1400 W/m2. The biasing conditions are p–n (solid green line; VG1 = −40 V, VG2 = 40 V), n–p (solid blue line; VG1 = 40 V, VG2 = −40 V), and p–p (dash blue line; VG1 = VG2 = 40 V). The red dashed rectangle in the main panel shows the maximum power conversion efficiency area (Pel). Top inset: Schematic of WSe2 monolayer device with split gate electrodes. Lower inset: electrical power (Pel) versus voltage under incident illumination. Reproduced with permission from [157], Nature Publishing Group, 2014.
Electronics 06 00043 g013
Figure 14. (a) Schematic of a top-gated MoS2 FET and the optical setup. (b) Absorption (Abs), EL, and PL spectra on the same 1L-MoS2. Reproduced with permission from [202], American Chemical Society, 2013.
Figure 14. (a) Schematic of a top-gated MoS2 FET and the optical setup. (b) Absorption (Abs), EL, and PL spectra on the same 1L-MoS2. Reproduced with permission from [202], American Chemical Society, 2013.
Electronics 06 00043 g014
Figure 15. Performance of a monolayer MoS2 phototransistor. (a) Ids–Vds characteristic of the device in the dark and under different illumination intensities. (b) Photoresponsivity of a similar monolayer MoS2 device as a function of illumination wavelength. Inset: Three-dimensional schematic view of the monolayer MoS2 photodetector and the focused laser beam used to probe the device. Reproduced with permission from [209], Nature Publishing Group, 2013.
Figure 15. Performance of a monolayer MoS2 phototransistor. (a) Ids–Vds characteristic of the device in the dark and under different illumination intensities. (b) Photoresponsivity of a similar monolayer MoS2 device as a function of illumination wavelength. Inset: Three-dimensional schematic view of the monolayer MoS2 photodetector and the focused laser beam used to probe the device. Reproduced with permission from [209], Nature Publishing Group, 2013.
Electronics 06 00043 g015
Figure 16. (a) The schematic band diagrams of ITO (gate)/Al2O3 (dielectric)/single-, double-, and triple-layer MoS2 (n-channel) under the gate bias for carrier (electron–hole) accumulation. (b) Photon energy-dependent (ΔQeff) plots indicate the approximate optical energy gaps to be 1.35 eV, 1.65 eV, and 1.82 eV for trilayer, bilayer, and monolayer MoS2 nanosheets, respectively. Reproduced with permission from [198], American Chemical Society, 2012.
Figure 16. (a) The schematic band diagrams of ITO (gate)/Al2O3 (dielectric)/single-, double-, and triple-layer MoS2 (n-channel) under the gate bias for carrier (electron–hole) accumulation. (b) Photon energy-dependent (ΔQeff) plots indicate the approximate optical energy gaps to be 1.35 eV, 1.65 eV, and 1.82 eV for trilayer, bilayer, and monolayer MoS2 nanosheets, respectively. Reproduced with permission from [198], American Chemical Society, 2012.
Electronics 06 00043 g016
Figure 17. (a) Illustration of PPC architecture, where the electric-field profile (in-plane, x-y) of the fundamental cavity mode (pristine cavity before WSe2 transfer) is embedded as the color plot. Reproduced with permission from [240], Nature Publishing Group, 2015. (b) Schematic of the DBR microcavity structure, a monolayer of MoS2, sandwiched between two SiO2 layers to form the cavity layer. Reproduced with permission from [239], Nature Publishing Group, 2014. (c) PL spectrum of the laser device with an excitation power of 30 mW at room temperature. Inset: Schematic configuration of the coupled SiO2 microsphere/microdisk optical cavity with the incorporation of MoS2 flake. Reproduced with permission from [235], American Chemical Society, 2015. (d) PL spectrum taken at 10 K when the pump intensity is above lasing threshold, showing whispering gallery modes (WGMs) at 612.2 nm, 633.7 nm, 657.6 nm, and 683.7 nm. Inset: Schematic image of the Si3N4/WS2/HSQ micro sphere sandwich structure. Reproduced with permission from [244], Nature Publishing Group, 2015.
Figure 17. (a) Illustration of PPC architecture, where the electric-field profile (in-plane, x-y) of the fundamental cavity mode (pristine cavity before WSe2 transfer) is embedded as the color plot. Reproduced with permission from [240], Nature Publishing Group, 2015. (b) Schematic of the DBR microcavity structure, a monolayer of MoS2, sandwiched between two SiO2 layers to form the cavity layer. Reproduced with permission from [239], Nature Publishing Group, 2014. (c) PL spectrum of the laser device with an excitation power of 30 mW at room temperature. Inset: Schematic configuration of the coupled SiO2 microsphere/microdisk optical cavity with the incorporation of MoS2 flake. Reproduced with permission from [235], American Chemical Society, 2015. (d) PL spectrum taken at 10 K when the pump intensity is above lasing threshold, showing whispering gallery modes (WGMs) at 612.2 nm, 633.7 nm, 657.6 nm, and 683.7 nm. Inset: Schematic image of the Si3N4/WS2/HSQ micro sphere sandwich structure. Reproduced with permission from [244], Nature Publishing Group, 2015.
Electronics 06 00043 g017
Figure 18. (a) Vertical cross section of the MoS2 amplifier device. (b) Schematic drawing of integrated amplifier in common-source configuration. (c) Transfer characteristic of the integrated amplifier realized with two transistors on the same MoS2 flake. (d) Voltage gain dependence on frequency of the small input signal. Reproduced with permission from [102], American Institute of Physics, 2012.
Figure 18. (a) Vertical cross section of the MoS2 amplifier device. (b) Schematic drawing of integrated amplifier in common-source configuration. (c) Transfer characteristic of the integrated amplifier realized with two transistors on the same MoS2 flake. (d) Voltage gain dependence on frequency of the small input signal. Reproduced with permission from [102], American Institute of Physics, 2012.
Electronics 06 00043 g018
Figure 19. Demonstration of an NOR-gate logic circuit based on monolayer MoS2 transistors. The output voltage Vout is shown for four different combinations of input states (1,0), (1,1), (0,1), and (0,0). The output is in the high state only if both inputs are in the low state. Left inset: The circuit is formed by connecting two monolayer MoS2 transistors in parallel and using an external 1 MΩ resistor as a load. Reproduced with permission from [97], American Chemical Society, 2011.
Figure 19. Demonstration of an NOR-gate logic circuit based on monolayer MoS2 transistors. The output voltage Vout is shown for four different combinations of input states (1,0), (1,1), (0,1), and (0,0). The output is in the high state only if both inputs are in the low state. Left inset: The circuit is formed by connecting two monolayer MoS2 transistors in parallel and using an external 1 MΩ resistor as a load. Reproduced with permission from [97], American Chemical Society, 2011.
Electronics 06 00043 g019
Figure 20. Demonstration of an integrated NAND logic gate and a static random-access memory (SRAM) cell on bilayer MoS2. (A) Optical micrograph of the NAND gate and the SRAM fabricated on the same bilayer MoS2 thin film. The corresponding schematics of the electronic circuits for the NAND gate and SRAM are also shown. (B) Output voltage of the flip-flop memory cell (SRAM). A Logic State 1 (or 0) at the input voltage can set the output voltage to Logic State 0 (or 1). In addition, the output logic state stays at 0 or 1 after the switch to the input has been opened. (C) Output voltage of the NAND gate for four different input states: (0,0), (0,1), (1,0), and (1,1). A low voltage below 0.5 V represents Logic State 0 and a voltage close to 2 V represents Logic State 1. Reproduced with permission from [249], American Chemical Society, 2012.
Figure 20. Demonstration of an integrated NAND logic gate and a static random-access memory (SRAM) cell on bilayer MoS2. (A) Optical micrograph of the NAND gate and the SRAM fabricated on the same bilayer MoS2 thin film. The corresponding schematics of the electronic circuits for the NAND gate and SRAM are also shown. (B) Output voltage of the flip-flop memory cell (SRAM). A Logic State 1 (or 0) at the input voltage can set the output voltage to Logic State 0 (or 1). In addition, the output logic state stays at 0 or 1 after the switch to the input has been opened. (C) Output voltage of the NAND gate for four different input states: (0,0), (0,1), (1,0), and (1,1). A low voltage below 0.5 V represents Logic State 0 and a voltage close to 2 V represents Logic State 1. Reproduced with permission from [249], American Chemical Society, 2012.
Electronics 06 00043 g020
Figure 21. A five-stage ring oscillator based on bilayer MoS2. (A) Optical micrograph of the ring oscillator constructed on a bilayer MoS2 thin film. (B) Schematic of the electronic circuit of the five-stage ring oscillator. The first five inverter stages form the positive feedback loop, which leads to the oscillation in the circuit. The last inverter serves as the synthesis stage. (C) Output voltage as a function of time for the ring oscillator at Vdd = 2 V. (D) The power spectrum of the output signal as a function of Vdd. From left to right, Vdd = 1.15 V and 1.2 to 2.0 V in steps of 0.1 V. The corresponding fundamental oscillation frequency increases from 0.52 to 1.6 MHz. Reproduced with permission from [249], American Chemical Society, 2012.
Figure 21. A five-stage ring oscillator based on bilayer MoS2. (A) Optical micrograph of the ring oscillator constructed on a bilayer MoS2 thin film. (B) Schematic of the electronic circuit of the five-stage ring oscillator. The first five inverter stages form the positive feedback loop, which leads to the oscillation in the circuit. The last inverter serves as the synthesis stage. (C) Output voltage as a function of time for the ring oscillator at Vdd = 2 V. (D) The power spectrum of the output signal as a function of Vdd. From left to right, Vdd = 1.15 V and 1.2 to 2.0 V in steps of 0.1 V. The corresponding fundamental oscillation frequency increases from 0.52 to 1.6 MHz. Reproduced with permission from [249], American Chemical Society, 2012.
Electronics 06 00043 g021
Figure 22. MoS2/graphene heterostructure memory device layout. (a) Three-dimensional schematic view of the memory device based on single-layer MoS2. Lower is the schematics of a heterostructure memory cell with a single-layer MoS2 semiconducting channel, graphene contacts and MLG floating gate. The MLG floating gate is separated from the channel by a thin tunneling oxide (1 nm Al2O3 + 6 nm HfO2) and from the control gate by a thicker blocking oxide (1 nm Al2O3 + 30 nm HfO2). (b) Simplified band di+agram of the Cr/HfO2/Gr/HfO2/MoS2 heterostructure in program and erase operation. (c) Temporal evolution of drain–source currents (Ids) in the erase (ON) and program (OFF) states. The curves are acquired independently for the program (10−8–10−7 A) and erase (10−12–10−10 A) current states and plotted on a common time scale. The drain–source bias voltage (Vds) is 50 mV, and the duration of the control-gate voltage (Vcg) pulse is 3s. Reproduced with permission from [252], American Chemical Society, 2013.
Figure 22. MoS2/graphene heterostructure memory device layout. (a) Three-dimensional schematic view of the memory device based on single-layer MoS2. Lower is the schematics of a heterostructure memory cell with a single-layer MoS2 semiconducting channel, graphene contacts and MLG floating gate. The MLG floating gate is separated from the channel by a thin tunneling oxide (1 nm Al2O3 + 6 nm HfO2) and from the control gate by a thicker blocking oxide (1 nm Al2O3 + 30 nm HfO2). (b) Simplified band di+agram of the Cr/HfO2/Gr/HfO2/MoS2 heterostructure in program and erase operation. (c) Temporal evolution of drain–source currents (Ids) in the erase (ON) and program (OFF) states. The curves are acquired independently for the program (10−8–10−7 A) and erase (10−12–10−10 A) current states and plotted on a common time scale. The drain–source bias voltage (Vds) is 50 mV, and the duration of the control-gate voltage (Vcg) pulse is 3s. Reproduced with permission from [252], American Chemical Society, 2013.
Electronics 06 00043 g022
Figure 23. “Optical erase-electrical write” and “electrical erase-optical write” operation of MoS2-PZT memories. Reproduced with permission from [250], American Chemical Society, 2015.
Figure 23. “Optical erase-electrical write” and “electrical erase-optical write” operation of MoS2-PZT memories. Reproduced with permission from [250], American Chemical Society, 2015.
Electronics 06 00043 g023
Table 1. Summary of semiconductor TMDCs materials (MX2) and the nature of their band gaps for monolayer (1L) to bulk samples. The direct (D) and indirect (I) bandgaps are as denoted. These values are extracted from both theoretical (T) and experimental (E) study. All data are referring to optical bandgaps at room temperature without considering quasiparticle (e.g., excitons) effects.
Table 1. Summary of semiconductor TMDCs materials (MX2) and the nature of their band gaps for monolayer (1L) to bulk samples. The direct (D) and indirect (I) bandgaps are as denoted. These values are extracted from both theoretical (T) and experimental (E) study. All data are referring to optical bandgaps at room temperature without considering quasiparticle (e.g., excitons) effects.
M or XSSeTe
Mo1L: ~1.8–1.9 eV (D)
Bulk: 1.2 eV (I) [10]
1L: 1.34 eV (D)(T) [11]
1L: 1.58 eV (D)(E) [12]
Bulk: 1.1 eV (I)(E) [13]
Bulk: 1.1 eV (I)(T) [14]
1L: 1.07 eV (D)(T) [15]
1L: 1.1 eV (D)(E) [16]
Bulk: 1.0 eV (I)(E) [17]
0.82 eV (I)(T) [18]
W1L: 1.94 eV (D)(T) [19]
1L: 2.14 eV (D)(E) [20]
Bulk: 1.35 eV (I)(E) [21]
1L: 1.74 eV (D)(T) [19]
1L: 1.65 eV (D)(E) [22]
Bulk: 1.1 eV (I)(D) [23]
Bulk: 1.2 eV (I)(E) [21]
1L: 1.14 eV (D)(T) [19]
Bulk: 0.7 eV (I)(T) [23]
Re1L: 1.43 eV (D)(T) [24]
1L: 1.55 eV (D)(E) [24]
Bulk: 1.35 eV (D)(T) [24]
Bulk: 1.47 eV (D)(E) [25]
1L: 1.34 eV (I)(T) [26]
1L: 1.239 eV (D)(T) [27]
1L: 1.47 eV (I)(E) [28]
2L: 1.165 eV (D)(T) [27]
2L: 1.32 eV (I)(E) [29]
4L: 1.092 eV (D)(T) [27]
Bulk: 1.06 eV (I)(T) [26]
Bulk: 1.18 (I)(E) [30]
Table 2. Performance of FETs based on multilayer TMDCs prepared by mechanical exfoliation and by chemical vapor deposition (as noted with CVD). Electron mobility is compiled here (or denoted by e). The hole mobility is denoted by h.
Table 2. Performance of FETs based on multilayer TMDCs prepared by mechanical exfoliation and by chemical vapor deposition (as noted with CVD). Electron mobility is compiled here (or denoted by e). The hole mobility is denoted by h.
MaterialsConfiguration
(Method)
Mobility (cm2·V−1·s−1)On/Off RatioSubthreshold Swing (mV·dec−1)Temperature (K)Reference
MoS2Back-gated3 300[41]
Back-gated401051000300[42]
Back-gated
(CVD)
2 × 10−2104 300[45]
Back-gated10010680300[46]
Back-gated2.4107 300[47]
Dual-gated517108140300[48]
Back-gated700 300[49]
Back-gated
(CVD)
174 × 108(bi-) 104(multi-) 300[50]
Four-terminal Back-gated306.5106 300[51]
Four-terminal Top-gated470(e)/480(h) 300[52]
ZrO2 & CNT Back-gated 106 3[53]
MoSe2Back-gated50106 300[54]
Four-terminal Back-gated on SiO2/parylene-C
(CVD)
50(Si) 160(parylene-C) 500106(e) 103(h) 295/100[55]
Four-terminal Back-gated
(CVD)
200(e)/150(h)106 275[56]
Back-gated
(CVD)
10103 300[57]
MoTe2Back-gated0.03(e)/0.3(h)2 × 103 300[58]
Four-terminal Back-gated20(h)105140300[59]
Ionic Liquid Top-gated30(e)/10(h) 140(e)/125(h)300[60]
Back-gated6(h)105 300[61]
Solid Polymer Electrolyte Back-gated7(e)/26(h)10590300[62]
Back-gated25.2(e)/1.5(h)2.1 × 105(e)/5.7 × 104(h) 280[63]
Back-gated2.04(h) 300[64]
WS2Iodine-transport Back-gated
(PVD)
105 300[65]
Ionic Liquid Top-gated20(e)/90(h) 90300[66]
Ionic Liquid Top-gated19(e)/12(h)10663(e)/67(h)300[20]
Back-gated234108 300[67]
Four-terminal Back-gated2010670300[68]
WSe2Four-terminal Top-gated500(h) 300[40]
Ionic Liquid Top-gated200(e/h) 330(e)/270(h)10 170/160/77[69]
Back-gated
(CVD)
350(h)108 300[70]
Back-gated
(CVD)
650(h)106250/140150/300/105[71]
Back-gated
(CVD)
10(h)104 300[72]
Back-gated92(h)~10 300[73]
ReS2Dual-gated12/26105148300/77[74]
Dual-gated1/5106750300/120[75]
Back-gated15.4107100300[76]
Back-gated1.5105 300[77]
Back-gated113 × 105 300[78]
Back-gated
(CVD)
7.2 × 10−2103 300[79]
ReSe2Top-gated0.1 300[27]
Back-gated6.71051300300[80]
Back-gated
(CVD)
1.36 × 10−3(h) 300[81]
Table 3. Performance of FETs based on monolayer, exfoliated TMDCs, and CVD TMDCs. Electron mobility is compiled here (or denoted by e). The hole mobility is denoted by h.
Table 3. Performance of FETs based on monolayer, exfoliated TMDCs, and CVD TMDCs. Electron mobility is compiled here (or denoted by e). The hole mobility is denoted by h.
MaterialsConfiguration
(Method)
Mobility (cm2·V−1·s−1)On/Off RatioSubthreshold Swing (mV·dec−1)Temperature (K)Reference
MoS2Top-gated21710874300[82]
Simulation Top-gated350101060300[96]
Back-gated12 300[83]
Top-gated320106 300[97]
Back-gated
(CVD)
0.04 300[98]
Back-gated10106 300[99]
Polymer Electrolyte Top-gated1010660300[100]
Ferro-electric polymer Top-gated220105300300[101]
Top-gated380106500300[102]
Top-gated1090108178300[103]
Back-gated
(CVD)
8107 300[84]
Multi-terminal Back-gated1000106 4[104]
Back-gated
(CVD)
6105 300[105]
Top-gated On SiO2/Si3N4
(CVD)
55(Si3N4)
24(SiO2)
107(SiO2) 300[106]
Top-gated
(CVD)
42.3106 300[107]
Dual-gated
(CVD)
190108170300[108]
MoSe2Back-gated
(CVD)
0.02(e)/0.01(h)102 300[109]
Back-gated
(CVD)
50106 300[110]
Back-gated
(CVD)
23(e)/17(h)105 300[111]
WS2Ionic Liquid Top-gated44(e)/43(h)10552(e)/57(h)300[20]
Back-gated50/140106 300/83[112]
Back-gated
(CVD)
4.1105 300[113]
Back-gated83/337 300/25[114]
WSe2Top-gated250(h)10660300[115]
Ionic Liquid Top-gated90(e)/7(h)104(e)
105(h)
300[116]
Polymer electrolyte Back-gated30(e)/180(h) 300[117]
Back-gated
(CVD)
100(h)108 300[70]
Table 4. Summary of transistor performances with different vertical heterostructures. Graphene and boron nitride nanosheets are represented as Gr and hBN, respectively. APTES: (3-aminopropyl) triethoxysilane.
Table 4. Summary of transistor performances with different vertical heterostructures. Graphene and boron nitride nanosheets are represented as Gr and hBN, respectively. APTES: (3-aminopropyl) triethoxysilane.
Materials and StructureOn/Off RatioOn-Current Density (A/cm2)Off-Current Density (A/cm2)Source–Drain Voltage (V)Reference
Gr/MoS2/Ti60 (290 K)
104 (150 K)
600.50.1[165]
hBN/Gr/WS2/Gr/Au1062000.0010.2[166]
Gr/MoS2/Ti/Au102150<10.01[169]
Gr/MoS2/Ti/Au103500012.57[170]
Gr/MoS2/Ti/Au1055 × 1040.10.5[171]
Gr/MoS2/Ti/Au3501.50.05[172]
Gr/MoSe2/Ti/Au10510000.050.5[167]
Au/Gr/MoS210610000.00140.5[173]
Gr/WSe2/Provskite/Gr106<1<10−5−1[174]
Gr/MoS2/Au10430000.30.5[168]
Gr/WSe2/Au10431000.310.5
APTES/Gr/WSe2/Pt3 × 104 (300 K)
5 × 107 (180 K)
110.0010.1[175]
Table 5. Performance of light-emitting diodes (LEDs) with various transition metal dichalcogenide (TMDC) structures.
Table 5. Performance of light-emitting diodes (LEDs) with various transition metal dichalcogenide (TMDC) structures.
Materials and StructureTypeEL Peak (nm)EL EfficiencyReference
Au/Cr/MoS2/Cr/AuSchottky junction6850.001%[202]
Ti/WSe2/Tip–i–n junction7400.06%[203]
Au/Pd/WSe2/Ti/Aup–n junction8000.1%[157]
Au/WSe2/Ptp–n junction7520.2%[158]
Au/WSe2/Aup–n junction7500.1%[159]
Ni/MoS2/WSe2/Aup–n junction80012%[184]
Gr/hBN/MoS2/hBN/GrTunnel junction6788.4%[204]
Gr/hBN/WS2/hBN/Gr6201.32%
Gr/hBN/WSe2/hBN/Gr7385.4%
Table 6. Performance of memory devices based on TMDCs and other materials.
Table 6. Performance of memory devices based on TMDCs and other materials.
MaterialsP/E RatioEndurance (Cycles)Retention (s)Operating Voltage(V)Reference
MoS2 (with PZT)5 × 103 1031[101]
1045001040.1[250]
1041202 × 1030.05[251]
Graphene/MoS21041202 × 1030.05[252]
Graphene/hBN/MoS21051201.4 × 1030.05[253]
1051051048[254]
MoS2/hBN/BP50401030.05[113]

Share and Cite

MDPI and ACS Style

Ye, M.; Zhang, D.; Yap, Y.K. Recent Advances in Electronic and Optoelectronic Devices Based on Two-Dimensional Transition Metal Dichalcogenides. Electronics 2017, 6, 43. https://doi.org/10.3390/electronics6020043

AMA Style

Ye M, Zhang D, Yap YK. Recent Advances in Electronic and Optoelectronic Devices Based on Two-Dimensional Transition Metal Dichalcogenides. Electronics. 2017; 6(2):43. https://doi.org/10.3390/electronics6020043

Chicago/Turabian Style

Ye, Mingxiao, Dongyan Zhang, and Yoke Khin Yap. 2017. "Recent Advances in Electronic and Optoelectronic Devices Based on Two-Dimensional Transition Metal Dichalcogenides" Electronics 6, no. 2: 43. https://doi.org/10.3390/electronics6020043

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop