Expand this Topic clickable element to expand a topic
Skip to content
Optica Publishing Group

Review of photo response in semiconductor transition metal dichalcogenides based photosensitive devices

Open Access Open Access

Abstract

Various performance aspects and the photocurrent generation mechanisms of photosensitive devices based on two-dimensional transition metal dichalcogenides are reviewed. The work begins with discussions on light absorption and charge separation mechanisms in TMD-based photodetectors. Then, performances of simple metal-TMDs-metal detectors are discussed in terms of responsivity and response time. Thereafter, novel photosensitive devices involving TMDs, either incorporating other materials or novel photosensitive device structure designs, are summarized, and basic ideas behind these devices to improve the photodetection performance are elaborated. Lastly, two valley degree of freedom related photoresponses, circular polarized photogalvanic effect and valley Hall effect, are presented. The review ends with prospectives of future challenges and opportunities in developing TMDs based photodetection devices.

© 2016 Optical Society of America

Corrections

17 June 2016: Corrections were made to the title.

1. Introduction

The conversion of absorbed photons into electric signal and conversion of injected electric current into light emission are the two faces of a coin that governs the major principles of optoelectronic applications. The former mainly leads to light detection related application such as video imaging, biomedical imaging, night vision, gas sensing and motion detection; while the latter mainly corresponds to the light emission application such as light emission diode (LED) and lasers. Although all these optoelectronics devices have already been commercialized and accompanied our everyday life for decades, the need for better performance optoelectronic devices keeps growing. In terms of photodetection which is the major scope of this review, ultimate performance photodetection platforms in terms of operation speed, quantum efficiency, spectrum range, flexibility, transparency and CMOS integrability are highly desired.

The emergency of two dimensional (2D) materials [1] with distinct physical, chemical and mechanical properties promise their great potentials for photodetection [2–5]. Compared to three dimensional semiconductors, two dimensional materials offer additional advantages in context of photodetection because of their transparency, mechanical flexibility, electronically highly tunable and easy processing. These advantages great facilitate applications in flexible optoelectronics and large scale CMOS integration into optoelectronics networks. A variety of prototype photodetection devices based on 2D materials were demonstrated during the last decade. These effort started with graphene, a pioneer of 2D family, showing many fantastic capabilities, such as broadband detection from ultraviolet (UV) to Terahertz (THz), ultrafast response time [6–8] and tunable optical properties via electrostatic doping [9, 10]. Many of these unique capabilities have already sparked enormous interests for commercialization. Comparing with graphene based photodetectors, 2D transitional metal dichalcogenides (2D-TMDs) processes sizable energy gap, allows lower dark currents and higher responsivity, but the operation speed and response spectrum range is limited. Although the current performance of TMDs based photodetectors cannot compete with the silicon based technology for conventional applications, the large excitonic effect and novel valley degree of freedom of TMDs enables unconventional applications that is not possible with conventional materials and (semi)metallic 2D counterpart such as graphene.

Despite its not until the recent few years the research of TMDs revives, several reviews have been available on optoelectronics and nanophotonics of 2D-TMDs [2–5] (TMDs for short below), in this work we mainly focus on the current state-of-the-art of phototectors and scanning photocurrent studies of photoresponse based on 2H-phase TMDs. First, the various performance aspects and photocurrent generation mechanism of typical metal-TMDs-metal photodetectors made by single species of TMDs are compared and summarized. Second, we discuss various approaches and their mechanisms in improving the performance over single species TMDs based photodetectors. Then, two valley degree of freedom related photoresponses of TMDs: circular polarized galvanic effect and valley Hall effect, are discussed. At last, we conclude with the major challenges and future opportunities in this field of research.

2. Performance of TMDs based photosensing device

The photodetection and many other optoelectronics applications relies on the light absorption and subsequent conversion of absorbed photons into an electric current. From this point of view, the optical absorption spectrum of TMDs and the electric current generation mechanism from absorbed photons are the two key components toward understanding the photoresponse of the TMDs.

2.1 Spectrum response of TMDs and photocurrent generation mechanisms

2H phase TMDs consist of a single layer of transition metal atoms sandwiched between two layers of chalcogen atoms in the trigonal prismatic structure as shown in Fig. 1(a), this lattice structure have out-of-plane mirror symmetry and broken in-plane inversion symmetry. Bulk TMDs are indirect-bandgap semiconductors with conduction band minimum and valence band maximum located at the Q and the Γ points, while in the monolayer limit, direct-bandgap semiconductors with gaps located at the K and the K’ points is formed [11, 12], this results in strong photoluminescence in monolayer TMDs [13, 14]. The light absorption capability of monolayer TMDs is dominated by direct transition between valence and conduction band states (band-to-band transition) around the K and K’ points, these transitions are strongly modified by the electron hole interaction due to reduced dielectric screening effect relative to the bulk. The exciton binding energy in monolayer TMDs ranges from 0.5 to 0.9 eV according to theoretical calculations [15–17] and optical/scanning tunneling spectroscopy measurement [18–22]. Higher order excitonic quasiparticles such as trions and biexcitons are also reported in monolayer TMDs [23–26]. These high order excitonic states can survive at room temperature in monolayer TMDs due to relatively large binding energy comparing to conventional quasi-2D semiconductors. These have important consequences for optical absorption spectrum of monolayer TMDs as significant oscillation strength is transferred from band-to-band transition to resonant absorption of excitonic states. In addition to that, the strong excitonic effect also leads to relatively short radiative lifetime [27] which may put the fundamental limit of the operation speed of a photo sensing device [28].

 figure: Fig. 1

Fig. 1 (a). Crystal structure of TMD. (b). The honeycomb lattice structure and first Brillouin zone with high-symmetry points. (c).The optical selection rules in MoS2, the left circular polarized light/right circular polarized light can selected to excite carriers in K/K' valley respectively. (d). The absorption spectra and PL spectra of one-layer and two-layer MoS2, the spectra are displaced along the vertical axis for clarity.[From Mak.K.F. et. al, Phys Rev Lett 105.136805(2010)] (e). Schematic representation of photocurrent generation in biased TMD devices. (f). Schematic representation of photocurrent generation due to photothermalelectric effect.

Download Full Size | PDF

The interaction between light and monolayer TMDs is very prominent, comparable to or even stronger than that of graphene in visible range. Figure 1(d) shows the absorption spectrum of a monolayer MoS2 [14]: below the A-exciton resonance, the absorbance is relatively low and negligible. The absorbance increases to 10% at photon energy in resonance with A/B excitons and keep climbing with photon energy due to band nesting effect of C band [29, 30], at 3 eV it can reach 30% [31]. A summary of basic electronic properties of different TMDs including the band masses, optical gaps, exciton binding energy conduction/valence band spin-orbital splitting can be found in a recent review [4]. The direct bandgaps in monolayer TMDs fall into visible and near infrared wavelength range, it not only depends on the TMDs species and number of layers but also can be continuously tuned with binary/ternary TMDs alloys [32–34].

Although strongly correlated to each other, the absorption spectrum not necessarily coincides with the photo response spectrum of a photo detection device as the current generation also relies on the efficient separation of charges that are created by absorption of photons. Two major charge separation mechanisms: photovoltaic effect (PV), which is based on the separation of photo generated charges by built-in electric fields at junctions; and photo-thermalelectric effect (PTE), which is based on the charge separation by thermalelectric power and electron temperature difference induced by the photoexciation, are both reported in TMDs based photodetectors [35–38], it was found that the relative contribution of PV and PTE can be tuned by the junction configurations and illumination photon energies [37, 38]. Thus both the excitation photo energy and the junction configuration are critical to the function mechanisms. With excitation in resonance with the excitonic state, despite the advances on the mechanism studies about charge separation, key issues for photodetectors because of extremely strong excitonic effects, such as the dissociation process of various excitonic states, are still not clear. Fundamental questions regarding these issues such as the exciton dissociation time and dissociation mechanism remain to be elucidated. On the other hand, the nature of the TMDs and metal contact, whether it’s Ohm contact and the height of the Schottky barriers, also strongly affect the charge separation and thus the response time and responsivity of the photodectors. It’s found the low Schottky barrier device exhibits a very high photo gain and but very slow response speed. In contrast, the high Schottky-contact barrier device display a fast response time, but the photo gain and specific detectivity decrease by several orders of magnitude [39].

2.2 Photo responsivity and response time

The reported photo responsivity and response time of TMDs based photodetectors spread over very large parameter ranges. Table 1 has summarized the major device performance parameters of metal-TMDs-metal structure photodectors reported so far. As shown in the table, the responsivities of different photodetectors range from 22 µA/W [40] to up to 880 A/W [41], while the response time ranges from few µs [42] to second [41] are reported in different devices by different groups. A rough “law” that can be drawn from Table 1 is that a device shows large responsivity turns to be slow in operation speed and vice versa. For example, photoconductive exfoliated monolayer-MoS2 detector exhibits high external responsivity of 880 A/W, but operates at very narrow bandwidth of 0.1 Hz [41]; while another fast device made from CVD grown monolayer MoS2 exhibits few ms response time, but the responsivity is only 3 mA/W [43]. Doping and trapping of photogenerated carriers by the impurity states has been identified as the main mechanism that change the device conductance in TMDs [42, 44]. This directly explains the sample dependent response time and responsivity reported in the literatures. Actually it’s not a surprise the performance of TMDs based photodectors are so different as reported from different groups or even from different works of the same group, considering that different TMDs materials (exfoliated or artificially grown), fabrication processes and electrode materials are used. It simply implies the performance of TMDs summarized in Table 1 should have considerable contribution from extrinsic effects related to impurity and defects states.

Tables Icon

Table 1. Comparison of Device Performance of TMD based Photodetectors

In terms of response speed, ultimate performance such as “ultrafast response” down to few ps [28] and persistent photoconductivity (PPC) over minutes [59] are both reported in TMDs. In the following, we'd like to discuss these two apparently conflicting circumstances in more details.

The standard method to measure the ultrafast transient response of photodetector is two-pulse photovoltage correlation (TPPC) [7]. This technique is performed by Wang et al. on a monolayer MoS2 photodector [28]. The as measured TPPC result (|ΔVc| as function of the time delay Δt between the two pulses) as shown in Fig. 2a exhibits transient photocurrent response with two distinct timescales: a fast component on the order of 3-5 ps, which is attributed by the Wang et al. to the short lifetime of the photoexcited carriers due to fast electron capture process by defects via Auger scattering (Fig. 2(b) II); a slow component on the order of 80-110 ps, which is attributed to carrier trap process by slow defects (Fig. 2(b) III). The fast timescale implies the ultrafast TMD photodetector can achieve current modulation bandwidths in 200-300 GHz range, but the modulation depth is limited by the relative contribution to the total photocurrent of this fast component. We note that the fast response time measured by Wang et al is not due to an improvement in the device. The captured electrons and holes could still be thermal excited and result in the very slow response after the transient response. The TPPC measurement can only resolve the fast response component of the photocurrent, the slow recovering part, ranging from tens of µs to several minutes as summarized in Table 1, is not magnified in a TPPC, but it may contribute to the major responsivity of the devices due to the accumulated response over relatively long time period after the transient response.

 figure: Fig. 2

Fig. 2 (a). TPPC (Vc) as a function of the time delay (Δt) between two pulses show a transient photocurrent response with two distinct timescales (i) a fast time scale of ~4.3ps and(ii) a slow timescale of ~105ps. [From Wang, H. N. et al., Nature Commu 6:8831(2015).] (b). The three temporal regions of ultrafast carrier dynamics of MoS2, (I) carriers thermalization and cooling process within 500 fs. (II) electrons and holes captured by fast defects within 1-2 ps (III) electrons captured by slow defects on a 60-70 ps timescale [From Wang, H. N. et al., Nano Lett. 15, 339−345(2015).] (c).The photoresponse of the MoS2 device can be classified into 5 stages. In addition to the photoresponse due to band-band transition (stage 2 and 4), the device exhibits a slow increase in the photocurrent under illumination (stage 3) and the PPC effect (stage 5). [From Wu, Yueh-Chun et al., Scientific Reports 5:11472(2015).] (d). The trapped electrons and holes are thermally excited to defects and result in PPC effect.

Download Full Size | PDF

Persistent photoconductivity (PPC) refers to sustained conductivity after photo illumination is terminated (Fig. 2(c)). From a general point of view, as photoresponse always takes finite time depending on the photocarrier dynamics and transport in the device, remaining conductivity should be universal in any photoresponse including ultrafast graphene based photodetector with several ps response time. Here we emphasize the PPC refers to sustained conductivity over relatively long time scale (on the order of seconds) after photo illumination. As the PPC effect last over long timescale, it may significantly modifies the transport properties, thus ideally the transport phenomena can be controlled through the history of the photon irradiation through PPC effect in optoelectronic devices.

Yueh-chun Wu et al. studied the PPC dependence on the temperature, the photon dose, the excitation photon energy and substrate effect in monolayer MoS2 field effect transistors systematically and conclude the effect can be attributed to random localized potential fluctuations (RLPF) in the devices [59]. The random localized potential fluctuation arise either from intrinsic disorders or extrinsic charge impurities which can localize the photogenated electron holes to local minima, thus hinder the recombination of photoexcited electron-hole pairs. After the photo illumination is off, the thermal excitation of carriers to higher energy states above the mobility edge produces photocurrent and results in the PPC effect, as shown in Fig. 2(d). Another two mechanisms: large lattice relaxation (LLR) and the microscopic carrier (MB) can be ruled out in photoresponse of TMDs as these two mechanisms are inconsistent with the experimentally observed temperature/excitation photon energy dependence and stretched exponential decay of the PPC effect respectively. Although the MB effect is not significant in photo response of pure TMD device, it is account for the PPC observed in various TMD based hybrid structures, such as MoS2/graphene [60, 61], quantum-dot/graphene [62], MoS2/Black Phosphorus [63], MoS2/Si [64] and chlorophyll/graphene heterostructures [65] as described in the following session.

2.3 Novel photosensitive device based on TMDs

So far, all discussions are based on photo sensing device built from individual TMDs species and concerns simple metal-TMDs-metal structure, which is a starting prototype device to study photocurrent response and understand the basic light-electric conversion mechanisms. However, frankly speaking, the performance of such device is not impressive comparing to traditional semiconductors and graphene.

Enormous efforts have been invested to enhance the performance of TMDs based photosensing devices by incorporating other materials especially two dimensional materials and designing novel device structure to enhance certain performance such as responsivity, spectrum range and response time. These new designs mainly fall into two categories, in the first category, novel device structures is designed to enhance certain performance: such as PN junctions is built to enhances the separation between electron and hole more efficiently; in the second category, heterostructures between TMD and other 2D materials or 3D semiconductors is employed to integrating advantage of different material systems.

PN junctions can facilitate separation of photoexcited electron-hole pairs and thus achieving better quantum efficiency. PN junctions on TMDs can be realized using either electrostatic doping [47, 66] or chemical doping [67, 68], as shown in Fig. 3(a). The responsivities of the photodetectors with a PN junction vary from several mA/W to several A/W depends on the working conditions of devices. A photoresponsivity of 5.07A/W with EQE of 1200% on 500 nm excitation photon energy was reported in MoS2 PN junction under a bias voltage of 1.5 V and a gate voltage of 0 V [67]. Plasmonic structure [69, 70], dielectric cavity [71], quantum dots [72, 73] or organic dye molecules [74] were incorporated into TMD devices to adjust the spectrum response of TMD and enhance the responsivity at given spectrum range, as shown in Fig. 3(b). The highest responsivity of 1.6*104 A/W was reported for photodectors based on the hybrid of MoS2 and graphene quantum dots [73].

 figure: Fig. 3

Fig. 3 (a). CPGE in bilayer WSe2 with excitation photon energy (1.17eV) that is below the indirect bandgap. [From Yuan, H. T. et al., Nat Nanotechnol 9, 851-857 (2014).] (b). The σ + and σ- excitations which are on-resonance with A exciton—a hole in the upper VB and electron in the CB lead to photocurrent generation in K and -K valleys, respectively. (c). CPGE current is generated due to interband transitions with excitation photon energy on-resonance with excitonic transition (1.96eV) and disappears with off-resonance excitations. [From Eginligil, M. et al., Nat Commun. 6 (2015).] (d). The valley-dependent optical selection rules and the electrons at the K and K′ valleys that possess opposite Berry curvatures (left) and a photoinduced AHE driven by a net valley polarization (middle) and an image of the Hall bar device (right). (e). The source drain bias dependence of the Hall voltage for the monolayer device. [From Mak, K. F. et al., Science 344, 1489-1492 (2014).]

Download Full Size | PDF

TMD has also been integrated into 3D semiconductors such as silicon to form a heterojunction in order to adjust the spectrum response (Fig. 3(d)) [64, 75–77]. A heterojunction with n-type MoS2 and P-type Si was reported to have a maximum responsivity of 300 mA/W and response speed of 3 µs with spectral response ranging from 350 to 1100nm [77]. While hybridding with 3D materials has the advantage of compatibility with current matured fabrication technologies, hybridding TMDs with other 2D materials, such as graphene [30, 61, 78–80], black phosphorus [63, 81] or another kind of TMDs [80, 82–85] provide more fantastic possibilities. Graphene-TMDs structures are the most studied structure of this kind and provides multiple advantages: first, graphene can be employed as contact material of TMDs, graphene has better optical transparency and minimal effect on band structure of TMDs compared to metal contact; second, the Dirac point of graphene usually lie in the bandgaps of TMDs, thus result in more efficient extraction of both electrons and holes from TMDs; third, a graphene-TMDs heterojunction will also benefit from the enhancement of light interaction with TMDs, the reported EQE of graphene-TMD heterojuctions is around 10%~55%; at last, the barrier height for graphene-TMD can be tuned by tuning the electric gating of Fermi level in graphene, and thus control the separation and transport of the excited electrons and holes with an external field [79]. Beyond graphene, incorporating TMD with BP or another kind of TMD will form different type of heterojunctions depending on the band alignment of the two materials. This allows for bandgap engineering of heterostructures, and the convenient electrostatically tuning of the carrier densities and band alignments of two-dimensional semiconductors also offers an alternative tuning knob for band engineering beyond switching materials. Such example is demonstrated by contacting layers of p-type WSe2 and n-type MoS2 to form a PN heterojunction (Fig. 3(c)) [82, 83]. Besides responsivity enhancement, the spectrum response can also be extent using these heterostructures. For example, an infrared (1.55 um) photodetector was reported using a type-II MoTe2/MoS2 van der waals heterostructures [85].

3. Valley related photoresponse

The broken inversion symmetry in monolayer TMDs yields an effective magnetic field (Berry curvatures Ω) in momentum space with opposite sign at the K and K’s valley [86, 87]. The berry curvature together with orbital magnetic moments gives rise to unique valley–dependent optical properties in monolayer TMDs. Direct addressing valley degree of freedom in layered TMDs with circular polarized light enables the opportunity to develop valley based information technology with TMDs and finally evokes the enormous research interests in TMDs [82, 86, 88, 89]. This is eight years after TMDs was first exfoliated into monolayer [1] and 3 year after monolayer MoS2 was found to be a direct bandgap semiconductor [13, 14]. The valley polarized photoluminescence after left/right circular polarized optical excitation is a direct way to test the valley degree of freedom in TMDs. On the other hand, helicity resolved photocurrent measurement of circular polarized photogalvanic effect(CPGE) and valley Hall effect (VHE) provide spin-valley related signature of photoresponse that can be measured experimentally.

3.1 Circular polarized photogalvanic effect

Benefiting from strongly couple spin-valley degree of freedom, CPGE has been demonstrated to generate spin-coupled valley polarization in TMDs, which enables additional degree of control for quantum-confined spintronics/valleytronics devices. CPGE in TMDs was first demonstrated by Yuan et al. in bilayer WSe2 with excitation photon energy (1.17eV) that is below the indirect bandgap. In this scheme, the current is generated due to intraband absorption at Λ valley within the conduction band (CB) as shown in Fig. 4(b). Under obliquely illumination with nonzero θ, the obtained photocurrent exhibits strong dependence on light circular polarization. The helicity dependent photocurrent arise from the asymmetric optical excitation of the splitting bands, the dependence of the photocurrent on light helicity indicates direct relationship between light helicity and spin or orbital angular momentum near WSe2 valleys. Furthermore, as the CPGE phenomenon is highly sensitive to degree of inversion asymmetry, applying a perpendicular electric field provides a convenient control of the CPGE current in bilayer WSe2 [90]. Later on, Eginligil et al. demonstrates the CPGE effect in inversion symmetry breaking chemical vapor deposition grown monolayer MoS2 in spin-valley coupled K valleys. In this work, CPGE current is generated due to interband transitions with excitation photon energy on-resonance with excitonic transition (1.96 eV, as shown in Fig. 4(c)), this CPGE effect disappears with off-resonance excitations. In both work, besides CPGE, linear polarized photogalvanic effect is also observed.

 figure: Fig. 4

Fig. 4 (a). The Schematic of a PN junction in monolayer WSe2 device controlled by two local gates. [From Baugher, Britton W. H. et al., Nat Nanotechnol 9, 262-267 (2014).] (b). The Schematic of a MoS2-GQDs heterostructure phototransistor [From Chen, C. Y. et al., Scientific Reports 5:11830 (2015).] (c).The Schematic of a MoS2/WSe2 heterostructure. [From Lee, Chul-Ho et al., Nat Nanotechnol 9,676-681 (2014).] (d). The Geometry of a MoS2/Si heterojunction diode. [From Lopez-Sanchez, Oriol et al., ACS Nano 8, 3042-3048, (2014).]

Download Full Size | PDF

3.2 Valley Hall effect

Although CPGE provides a signature of unique spin-valley degree of freedom of TMDs through photocurrent measurement, direct demonstration of so-called valley Hall effect gives more clear and direct signature of valley degree of freedom in TMDs. Mak et al. [91] experimentally observed the valley Hall effect (VHE) in a monolayer of MoS2. This is done in a hall bar device as shown in Fig. 4(d), when the hall bar is illuminated with circular polarized light, the carriers are excited into a specific valley depending on the excitation helicity of the light. This is because in a monolayer hall bar device with inversion symmetry broken, carriers in two valleys experience effective magnetic fields with opposite signs causing by the berry curvature. The opposite magnetic field drives electrons in different valleys to move in opposite directions perpendicular to the drift current as it biased in a hall bar device and gives the VHE.The tiny Hall voltage is observed by high frequency modulation between left/right circular polarization and detecting the VHE voltage modulation with a lock-in amplifier. In a bilayer, as inversion symmetry is recovered, VHE effect disappears, but it revokes with an perpendicular electric field added to break the inversion symmetry in the bilayer and makes the VHE effect controllable with a top gate, this is also observed experimentally in a recent work [92, 93].

4. Prospectives

Although TMDs has been demonstrated great potential for high performance photosensing with the possibility to incorporate fantastic spin-valley coupled degree of freedoms, many challenges remain to be overcome and new opportunities remain to be explored. With a lot of previous efforts devoted to TMDs based photodetection, our understanding of the intrinsic photoresponse of TMDs are still limited, further advance in understanding the function mechanisms relies on ultrafast dynamics studies of photoexcited carrier [94–111], especially excitonic quasiparticles in TMDs [26, 112–126], contact between metal/hybrid materials and TMDs [127–132] and better control of defects and impurity levels in material growth [50, 133–137] and device fabrication for studying the intrinsic photo response of TMDs. In terms of device performance, TMDs suffer from certain limitations such as limited spectrum response, and operation bandwidth et al., this will reply on innovation of the device structures and hybriding with other materials systems, especially van der waals hetrostructures with other 2D materials with different functionalities. Initial efforts toward spin- and valley-polarized light emission and detection carrying valley degree of information has been practiced [138, 139]. Incorporating the coupled valley-spin degree of freedom into TMDs based optoelectronic is an attractive endeavor and it can possibility renovate future information technology.

Acknowledgments

The authors would like to acknowledge the funding support from the following: the National Basic Research Program of China (973 Grant Nos. 2012CB921300, 2014CB920900), the National Natural Science Foundation of China (NSFC Grant Nos. 11274015), the Recruitment Program of Global Experts, Beijing Natural Science Foundation (Grant No. 4142024) and the Specialized Research Fund for the Doctoral Program of Higher Education of China (Grant No.20120001110066).

References and links

1. K. S. Novoselov, A. K. Geim, S. V. Morozov, D. Jiang, Y. Zhang, S. V. Dubonos, I. V. Grigorieva, and A. A. Firsov, “Electric field effect in atomically thin carbon films,” Science 306(5696), 666–669 (2004). [CrossRef]   [PubMed]  

2. Q. H. Wang, K. Kalantar-Zadeh, A. Kis, J. N. Coleman, and M. S. Strano, “Electronics and optoelectronics of two-dimensional transition metal dichalcogenides,” Nat. Nanotechnol. 7(11), 699–712 (2012). [CrossRef]   [PubMed]  

3. F. H. L. Koppens, T. Mueller, P. Avouris, A. C. Ferrari, M. S. Vitiello, and M. Polini, “Photodetectors based on graphene, other two-dimensional materials and hybrid systems,” Nat. Nanotechnol. 9(10), 780–793 (2014). [CrossRef]   [PubMed]  

4. K. F. Mak and J. Shan, “Photonics and optoelectronics of 2D semiconductor transition metal dichalcogenides,” Nat. Photonics 10(4), 216–226 (2016). [CrossRef]  

5. F. Xia, H. Wang, D. Xiao, M. Dubey, and A. Ramasubramaniam, “Two-dimensional material nanophotonics,” Nat. Photonics 8(12), 899–907 (2014). [CrossRef]  

6. F. Xia, T. Mueller, Y. M. Lin, A. Valdes-Garcia, and P. Avouris, “Ultrafast graphene photodetector,” Nat. Nanotechnol. 4(12), 839–843 (2009). [CrossRef]   [PubMed]  

7. D. Sun, G. Aivazian, A. M. Jones, J. S. Ross, W. Yao, D. Cobden, and X. Xu, “Ultrafast hot-carrier-dominated photocurrent in graphene,” Nat. Nanotechnol. 7(2), 114–118 (2012). [CrossRef]   [PubMed]  

8. K. J. Tielrooij, L. Piatkowski, M. Massicotte, A. Woessner, Q. Ma, Y. Lee, K. S. Myhro, C. N. Lau, P. Jarillo-Herrero, N. F. van Hulst, and F. H. L. Koppens, “Generation of photovoltage in graphene on a femtosecond timescale through efficient carrier heating,” Nat Nano 10(5), 437–443 (2015). [CrossRef]  

9. Z. Q. Li, E. A. Henriksen, Z. Jiang, Z. Hao, M. C. Martin, P. Kim, H. L. Stormer, and D. N. Basov, “Dirac charge dynamics in graphene by infrared spectroscopy,” Nat. Phys. 4(7), 532–535 (2008). [CrossRef]  

10. F. Wang, Y. Zhang, C. Tian, C. Girit, A. Zettl, M. Crommie, and Y. R. Shen, “Gate-variable optical transitions in graphene,” Science 320(5873), 206–209 (2008). [CrossRef]   [PubMed]  

11. S. Lebègue and O. Eriksson, “Electronic structure of two-dimensional crystals from ab initio theory,” Phys. Rev. B 79(11), 115409 (2009). [CrossRef]  

12. T. Cheiwchanchamnangij and W. R. L. Lambrecht, “Quasiparticle band structure calculation of monolayer, bilayer, and bulk MoS2,” Phys. Rev. B 85(20), 205302 (2012). [CrossRef]  

13. A. Splendiani, L. Sun, Y. Zhang, T. Li, J. Kim, C.-Y. Chim, G. Galli, and F. Wang, “Emerging photoluminescence in monolayer MoS2.,” Nano Lett. 10(4), 1271–1275 (2010). [CrossRef]   [PubMed]  

14. K. F. Mak, C. Lee, J. Hone, J. Shan, and T. F. Heinz, “Atomically thin MoS2: a new direct-gap semiconductor,” Phys. Rev. Lett. 105(13), 136805 (2010). [CrossRef]   [PubMed]  

15. P. Cudazzo, I. V. Tokatly, and A. Rubio, “Dielectric screening in two-dimensional insulators: Implications for excitonic and impurity states in graphane,” Phys. Rev. B 84(8), 085406 (2011). [CrossRef]  

16. A. Ramasubramaniam, “Large excitonic effects in monolayers of molybdenum and tungsten dichalcogenides,” Phys. Rev. B 86(11), 115409 (2012). [CrossRef]  

17. D. Y. Qiu, F. H. da Jornada, and S. G. Louie, “Optical spectrum of MoS2: many-body effects and diversity of exciton states,” Phys. Rev. Lett. 111(21), 216805 (2013). [CrossRef]   [PubMed]  

18. A. Chernikov, T. C. Berkelbach, H. M. Hill, A. Rigosi, Y. Li, O. B. Aslan, D. R. Reichman, M. S. Hybertsen, and T. F. Heinz, “Exciton binding energy and nonhydrogenic Rydberg series in monolayer WS2.,” Phys. Rev. Lett. 113(7), 076802 (2014). [CrossRef]   [PubMed]  

19. K. He, N. Kumar, L. Zhao, Z. Wang, K. F. Mak, H. Zhao, and J. Shan, “Tightly bound excitons in monolayer WSe2,” Phys. Rev. Lett. 113(2), 026803 (2014). [CrossRef]   [PubMed]  

20. Z. Ye, T. Cao, K. O’Brien, H. Zhu, X. Yin, Y. Wang, S. G. Louie, and X. Zhang, “Probing excitonic dark states in single-layer tungsten disulphide,” Nature 513(7517), 214–218 (2014). [CrossRef]   [PubMed]  

21. M. M. Ugeda, A. J. Bradley, S.-F. Shi, F. H. da Jornada, Y. Zhang, D. Y. Qiu, W. Ruan, S.-K. Mo, Z. Hussain, Z.-X. Shen, F. Wang, S. G. Louie, and M. F. Crommie, “Giant bandgap renormalization and excitonic effects in a monolayer transition metal dichalcogenide semiconductor,” Nat. Mater. 13(12), 1091–1095 (2014). [CrossRef]   [PubMed]  

22. C. Zhang, A. Johnson, C.-L. Hsu, L.-J. Li, and C.-K. Shih, “Direct imaging of band profile in single layer MoS2 on graphite: quasiparticle energy gap, metallic edge states, and edge band bending,” Nano Lett. 14(5), 2443–2447 (2014). [CrossRef]   [PubMed]  

23. K. F. Mak, K. He, C. Lee, G. H. Lee, J. Hone, T. F. Heinz, and J. Shan, “Tightly bound trions in monolayer MoS2.,” Nat. Mater. 12(3), 207–211 (2012). [CrossRef]   [PubMed]  

24. J. S. Ross, S. Wu, H. Yu, N. J. Ghimire, A. M. Jones, G. Aivazian, J. Yan, D. G. Mandrus, D. Xiao, W. Yao, and X. Xu, “Electrical control of neutral and charged excitons in a monolayer semiconductor,” Nat. Commun. 4, 1474 (2013). [CrossRef]   [PubMed]  

25. J. Shang, X. Shen, C. Cong, N. Peimyoo, B. Cao, M. Eginligil, and T. Yu, “Observation of excitonic fine structure in a 2D Transition-Metal Dichalcogenide Semiconductor,” ACS Nano 9(1), 647–655 (2015). [CrossRef]   [PubMed]  

26. Y. You, X.-X. Zhang, T. C. Berkelbach, M. S. Hybertsen, D. R. Reichman, and T. F. Heinz, “Observation of biexcitons in monolayer WSe2,” Nat. Phys. 11(6), 477–481 (2015). [CrossRef]  

27. J. Feldmann, G. Peter, E. O. Göbel, P. Dawson, K. Moore, C. Foxon, and R. J. Elliott, “Linewidth dependence of radiative exciton lifetimes in quantum wells,” Phys. Rev. Lett. 59(20), 2337–2340 (1987). [CrossRef]   [PubMed]  

28. H. Wang, C. Zhang, W. Chan, S. Tiwari, and F. Rana, “Ultrafast response of monolayer molybdenum disulfide photodetectors,” Nat. Commun. 6, 8831 (2015). [CrossRef]   [PubMed]  

29. A. Carvalho, R. M. Ribeiro, and A. H. Castro Neto, “Band nesting and the optical response of two-dimensional semiconducting transition metal dichalcogenides,” Phys. Rev. B 88(11), 115205 (2013). [CrossRef]  

30. L. Britnell, R. M. Ribeiro, A. Eckmann, R. Jalil, B. D. Belle, A. Mishchenko, Y. J. Kim, R. V. Gorbachev, T. Georgiou, S. V. Morozov, A. N. Grigorenko, A. K. Geim, C. Casiraghi, A. H. Castro Neto, and K. S. Novoselov, “Strong light-matter interactions in heterostructures of atomically thin films,” Science 340(6138), 1311–1314 (2013). [CrossRef]   [PubMed]  

31. L. Shang-Chun, M. Mohamed, and Z. Wenjuan, “Novel vertical hetero- and homo-junction tunnel field-effect transistors based on multi-layer 2D crystals,” 2d Mater. 3(1), 011010 (2016). [CrossRef]  

32. Y. Chen, J. Xi, D. O. Dumcenco, Z. Liu, K. Suenaga, D. Wang, Z. Shuai, Y.-S. Huang, and L. Xie, “Tunable band gap photoluminescence from atomically thin transition-metal dichalcogenide alloys,” ACS Nano 7(5), 4610–4616 (2013). [CrossRef]   [PubMed]  

33. S. Tongay, D. S. Narang, J. Kang, W. Fan, C. Ko, A. V. Luce, K. X. Wang, J. Suh, K. D. Patel, V. M. Pathak, J. Li, and J. Wu, “Two-dimensional semiconductor alloys: monolayer Mo1−xWxSe2,” Appl. Phys. Lett. 104(1), 012101 (2014). [CrossRef]  

34. L. M. Xie, “Two-dimensional transition metal dichalcogenide alloys: preparation, characterization and applications,” Nanoscale 7(44), 18392–18401 (2015). [CrossRef]   [PubMed]  

35. M. Buscema, M. Barkelid, V. Zwiller, H. S. J. van der Zant, G. A. Steele, and A. Castellanos-Gomez, “Large and tunable photothermoelectric effect in single-layer MoS2.,” Nano Lett. 13(2), 358–363 (2013). [CrossRef]   [PubMed]  

36. M. Fontana, T. Deppe, A. K. Boyd, M. Rinzan, A. Y. Liu, M. Paranjape, and P. Barbara, “Electron-hole transport and photovoltaic effect in gated MoS2 Schottky junctions,” Sci. Rep. 3, 1634 (2013). [CrossRef]   [PubMed]  

37. D. J. Groenendijk, M. Buscema, G. A. Steele, S. Michaelis de Vasconcellos, R. Bratschitsch, H. S. J. van der Zant, and A. Castellanos-Gomez, “Photovoltaic and photothermoelectric effect in a double-gated WSe2 device,” Nano Lett. 14(10), 5846–5852 (2014). [CrossRef]   [PubMed]  

38. Y. Zhang, H. Li, L. Wang, H. Wang, X. Xie, S.-L. Zhang, R. Liu, and Z.-J. Qiu, “Photothermoelectric and photovoltaic effects both present in MoS2.,” Sci. Rep. 5, 7938 (2015). [CrossRef]   [PubMed]  

39. W. Zhang, M.-H. Chiu, C.-H. Chen, W. Chen, L.-J. Li, and A. T. S. Wee, “Role of metal contacts in high-performance phototransistors based on WSe2 monolayers,” ACS Nano 8(8), 8653–8661 (2014). [CrossRef]   [PubMed]  

40. N. Perea-López, A. L. Elías, A. Berkdemir, A. Castro-Beltran, H. R. Gutiérrez, S. Feng, R. Lv, T. Hayashi, F. López-Urías, S. Ghosh, B. Muchharla, S. Talapatra, H. Terrones, and M. Terrones, “Photosensor device based on few-layered WS2 films,” Adv. Funct. Mater. 23(44), 5511–5517 (2013). [CrossRef]  

41. O. Lopez-Sanchez, D. Lembke, M. Kayci, A. Radenovic, and A. Kis, “Ultrasensitive photodetectors based on monolayer MoS2.,” Nat. Nanotechnol. 8(7), 497–501 (2013). [CrossRef]   [PubMed]  

42. D.-S. Tsai, K.-K. Liu, D.-H. Lien, M.-L. Tsai, C.-F. Kang, C.-A. Lin, L.-J. Li, and J.-H. He, “Few-layer MoS2 with high broadband photogain and fast optical switching for use in harsh environments,” ACS Nano 7(5), 3905–3911 (2013). [CrossRef]   [PubMed]  

43. Z. Yang, R. Grassi, M. Freitag, Y.-H. Lee, T. Low, and W. Zhu, “Spatial/temporal photocurrent and electronic transport in monolayer molybdenum disulfide grown by chemical vapor deposition,” Appl. Phys. Lett. 108(8), 083104 (2016). [CrossRef]  

44. M. M. Furchi, D. K. Polyushkin, A. Pospischil, and T. Mueller, “Mechanisms of photoconductivity in atomically thin MoS2.,” Nano Lett. 14(11), 6165–6170 (2014). [CrossRef]   [PubMed]  

45. Z. Yin, H. Li, H. Li, L. Jiang, Y. Shi, Y. Sun, G. Lu, Q. Zhang, X. Chen, and H. Zhang, “Single-Layer MoS2 Phototransistors,” ACS Nano 6(1), 74–80 (2012). [CrossRef]   [PubMed]  

46. W. Choi, M. Y. Cho, A. Konar, J. H. Lee, G. B. Cha, S. C. Hong, S. Kim, J. Kim, D. Jena, J. Joo, and S. Kim, “High-detectivity multilayer MoS2 phototransistors with spectral response from ultraviolet to infrared,” Adv. Mater. 24(43), 5832–5836 (2012). [CrossRef]   [PubMed]  

47. B. W. H. Baugher, H. O. H. Churchill, Y. Yang, and P. Jarillo-Herrero, “Optoelectronic devices based on electrically tunable p-n diodes in a monolayer dichalcogenide,” Nat. Nanotechnol. 9(4), 262–267 (2014). [CrossRef]   [PubMed]  

48. A. R. Klots, A. K. M. Newaz, B. Wang, D. Prasai, H. Krzyzanowska, J. Lin, D. Caudel, N. J. Ghimire, J. Yan, B. L. Ivanov, K. A. Velizhanin, A. Burger, D. G. Mandrus, N. H. Tolk, S. T. Pantelides, and K. I. Bolotin, “Probing excitonic states in suspended two-dimensional semiconductors by photocurrent spectroscopy,” Sci. Rep. 4, 6608 (2014). [CrossRef]   [PubMed]  

49. S. H. Lee, D. Lee, W. S. Hwang, E. Hwang, D. Jena, and W. J. Yoo, “High-performance photocurrent generation from two-dimensional WS2 field-effect transistors,” Appl. Phys. Lett. 104, 193113 (2014).

50. N. Perea-Lopez, Z. Lin, N. R. Pradhan, A. Iniguez-Rabago, A. L. Elias, A. McCreary, J. Lou, P. M. Ajayan, H. Terrones, L. Balicas, and M. Terrones, “CVD-grown monolayered MoS2 as an effective photosensor operating at low-voltage,” 2d Mater. 1, 011004 (2014).

51. M. Eginligil, B. Cao, Z. Wang, X. Shen, C. Cong, J. Shang, C. Soci, and T. Yu, “Dichroic spin-valley photocurrent in monolayer molybdenum disulphide,” Nat. Commun. 6, 7636 (2015). [CrossRef]   [PubMed]  

52. S. Ghosh, A. Winchester, B. Muchharla, M. Wasala, S. Feng, A. L. Elias, M. B. M. Krishna, T. Harada, C. Chin, K. Dani, S. Kar, M. Terrones, and S. Talapatra, “Ultrafast intrinsic photoresponse and direct evidence of sub-gap states in liquid phase exfoliated MoS2thin films,” Sci. Rep. 5, 11272 (2015). [CrossRef]   [PubMed]  

53. C. Jung, S. M. Kim, H. Moon, G. Han, J. Kwon, Y. K. Hong, I. Omkaram, Y. Yoon, S. Kim, and J. Park, “Highly crystalline CVD-grown multilayer MoSe2 thin film transistor for fast photodetector,” Sci. Rep. 5, 15313 (2015). [CrossRef]   [PubMed]  

54. N. R. Pradhan, J. Ludwig, Z. Lu, D. Rhodes, M. M. Bishop, K. Thirunavukkuarasu, S. A. McGill, D. Smirnov, and L. Balicas, “High photoresponsivity and short photoresponse times in few-layered WSe2 transistors,” ACS Appl. Mater. Interfaces 7(22), 12080–12088 (2015). [CrossRef]   [PubMed]  

55. H. Yamaguchi, J.-C. Blancon, R. Kappera, S. Lei, S. Najmaei, B. D. Mangum, G. Gupta, P. M. Ajayan, J. Lou, M. Chhowalla, J. J. Crochet, and A. D. Mohite, “Spatially resolved photoexcited charge-carrier dynamics in phase-engineered monolayer MoS2.,” ACS Nano 9(1), 840–849 (2015). [CrossRef]   [PubMed]  

56. J. Kwon, Y. K. Hong, G. Han, I. Omkaram, W. Choi, S. Kim, and Y. Yoon, “Giant photoamplification in indirect-bandgap multilayer MoS2 phototransistors with local bottom-gate structures,” Adv. Mater. 27(13), 2224–2230 (2015). [CrossRef]   [PubMed]  

57. L. Zeng, L. Tao, C. Tang, B. Zhou, H. Long, Y. Chai, S. P. Lau, and Y. H. Tsang, “High-responsivity UV-vis photodetector based on transferable ws2 film deposited by magnetron sputtering,” Sci. Rep. 6, 20343 (2016). [CrossRef]   [PubMed]  

58. A. S. Pawbake, R. G. Waykar, D. J. Late, and S. R. Jadkar, “Highly transparent wafer-scale synthesis of crystalline WS2 nanoparticle thin film for photodetector and humidity-sensing applications,” ACS Appl. Mater. Interfaces 8(5), 3359–3365 (2016). [CrossRef]   [PubMed]  

59. Y.-C. Wu, C.-H. Liu, S.-Y. Chen, F.-Y. Shih, P.-H. Ho, C.-W. Chen, C.-T. Liang, and W.-H. Wang, “Extrinsic origin of persistent photoconductivity in monolayer MoS2 field effect transistors,” Sci. Rep. 5, 11472 (2015). [CrossRef]   [PubMed]  

60. K. Roy, M. Padmanabhan, S. Goswami, T. P. Sai, G. Ramalingam, S. Raghavan, and A. Ghosh, “Graphene-MoS2 hybrid structures for multifunctional photoresponsive memory devices,” Nat. Nanotechnol. 8(11), 826–830 (2013). [CrossRef]   [PubMed]  

61. S. Rathi, I. Lee, D. Lim, J. Wang, Y. Ochiai, N. Aoki, K. Watanabe, T. Taniguchi, G.-H. Lee, Y.-J. Yu, P. Kim, and G.-H. Kim, “Tunable electrical and optical characteristics in monolayer graphene and few-layer MoS2 heterostructure devices,” Nano Lett. 15(8), 5017–5024 (2015). [CrossRef]   [PubMed]  

62. G. Konstantatos, M. Badioli, L. Gaudreau, J. Osmond, M. Bernechea, F. P. Garcia de Arquer, F. Gatti, and F. H. L. Koppens, “Hybrid graphene-quantum dot phototransistors with ultrahigh gain,” Nat. Nanotechnol. 7(6), 363–368 (2012). [CrossRef]   [PubMed]  

63. L. Ye, H. Li, Z. Chen, and J. Xu, “Near-infrared photodetector based on MoS2/black phosphorus heterojunction,” ACS Photonics 3(4), 692–699 (2016). [CrossRef]  

64. M. R. Esmaeili-Rad and S. Salahuddin, “High performance molybdenum disulfide amorphous silicon heterojunction photodetector,” Sci. Rep. 3, 2345 (2013). [CrossRef]   [PubMed]  

65. S.-Y. Chen, Y.-Y. Lu, F.-Y. Shih, P.-H. Ho, Y.-F. Chen, C.-W. Chen, Y.-T. Chen, and W.-H. Wang, “Biologically inspired graphene-chlorophyll phototransistors with high gain,” Carbon 63, 23–29 (2013). [CrossRef]  

66. J. S. Ross, P. Klement, A. M. Jones, N. J. Ghimire, J. Yan, D. G. Mandrus, T. Taniguchi, K. Watanabe, K. Kitamura, W. Yao, D. H. Cobden, and X. Xu, “Electrically tunable excitonic light-emitting diodes based on monolayer WSe2 p-n junctions,” Nat. Nanotechnol. 9(4), 268–272 (2014). [CrossRef]   [PubMed]  

67. M. S. Choi, D. Qu, D. Lee, X. Liu, K. Watanabe, T. Taniguchi, and W. J. Yoo, “Lateral MoS2 p-n junction formed by chemical doping for use in high-performance optoelectronics,” ACS Nano 8(9), 9332–9340 (2014). [CrossRef]   [PubMed]  

68. H.-M. Li, D. Lee, D. Qu, X. Liu, J. Ryu, A. Seabaugh, and W. J. Yoo, “Ultimate thin vertical p-n junction composed of two-dimensional layered molybdenum disulfide,” Nat. Commun. 6, 6564 (2015). [CrossRef]   [PubMed]  

69. J. D. Lin, H. Li, H. Zhang, and W. Chen, “Plasmonic enhancement of photocurrent in MoS2 field-effect-transistor,” Appl. Phys. Lett. 102(20), 203109 (2013). [CrossRef]  

70. T. Hong, B. Chamlagain, S. Hu, S. M. Weiss, Z. Zhou, and Y. Q. Xu, “Plasmonic hot electron induced photocurrent response at MoS2-metal junctions,” ACS Nano 9(5), 5357–5363 (2015). [CrossRef]   [PubMed]  

71. J. Zheng, R. A. Barton, and D. Englund, “Broadband coherent absorption in chirped-planar-dielectric cavities for 2D-material-based photovoltaics and photodetectors,” ACS Photonics 1(9), 768–774 (2014). [CrossRef]  

72. D. Kufer, I. Nikitskiy, T. Lasanta, G. Navickaite, F. H. L. Koppens, and G. Konstantatos, “Hybrid 2D-0D MoS2 -PbS quantum dot photodetectors,” Adv. Mater. 27(1), 176–180 (2015). [CrossRef]   [PubMed]  

73. C. Chen, H. Qiao, S. Lin, C. Man Luk, Y. Liu, Z. Xu, J. Song, Y. Xue, D. Li, J. Yuan, W. Yu, C. Pan, S. Ping Lau, and Q. Bao, “Highly responsive MoS2 photodetectors enhanced by graphene quantum dots,” Sci. Rep. 5, 11830 (2015). [CrossRef]   [PubMed]  

74. S. H. Yu, Y. Lee, S. K. Jang, J. Kang, J. Jeon, C. Lee, J. Y. Lee, H. Kim, E. Hwang, S. Lee, and J. H. Cho, “Dye-sensitized MoS2 photodetector with enhanced spectral photoresponse,” ACS Nano 8(8), 8285–8291 (2014). [CrossRef]   [PubMed]  

75. O. Lopez-Sanchez, E. Alarcon Llado, V. Koman, A. Fontcuberta i Morral, A. Radenovic, and A. Kis, “Light generation and harvesting in a van der Waals heterostructure,” ACS Nano 8(3), 3042–3048 (2014). [CrossRef]   [PubMed]  

76. B. Li, G. Shi, S. Lei, Y. He, W. Gao, Y. Gong, G. Ye, W. Zhou, K. Keyshar, J. Hao, P. Dong, L. Ge, J. Lou, J. Kono, R. Vajtai, and P. M. Ajayan, “3D band diagram and photoexcitation of 2D-3D semiconductor heterojunctions,” Nano Lett. 15(9), 5919–5925 (2015). [CrossRef]   [PubMed]  

77. L. Wang, J. S. Jie, Z. B. Shao, Q. Zhang, X. H. Zhang, Y. M. Wang, Z. Sun, and S. T. Lee, “MoS2/Si heterojunction with vertically standing layered structure for ultrafast, high-detectivity, self-driven visible-near infrared photodetectors,” Adv. Funct. Mater. 25(19), 2910–2919 (2015). [CrossRef]  

78. K. Roy, M. Padmanabhan, S. Goswami, T. P. Sai, G. Ramalingam, S. Raghavan, and A. Ghosh, “Graphene-MoS2 hybrid structures for multifunctional photoresponsive memory devices,” Nat. Nanotechnol. 8(11), 826–830 (2013). [CrossRef]   [PubMed]  

79. W. J. Yu, Y. Liu, H. Zhou, A. Yin, Z. Li, Y. Huang, and X. Duan, “Highly efficient gate-tunable photocurrent generation in vertical heterostructures of layered materials,” Nat. Nanotechnol. 8(12), 952–958 (2013). [CrossRef]   [PubMed]  

80. M. J. Park, J. K. Min, S. G. Yi, J. H. Kim, J. Oh, and K. H. Yoo, “Near-infrared photodetectors utilizing MoS2-based heterojunctions,” J. Appl. Phys. 118(4), 044504 (2015). [CrossRef]  

81. Y. Deng, Z. Luo, N. J. Conrad, H. Liu, Y. Gong, S. Najmaei, P. M. Ajayan, J. Lou, X. Xu, and P. D. Ye, “Black phosphorus-monolayer MoS2 van der Waals heterojunction p-n diode,” ACS Nano 8(8), 8292–8299 (2014). [CrossRef]   [PubMed]  

82. C.-H. Lee, G.-H. Lee, A. M. van der Zande, W. Chen, Y. Li, M. Han, X. Cui, G. Arefe, C. Nuckolls, T. F. Heinz, J. Guo, J. Hone, and P. Kim, “Atomically thin p-n junctions with van der Waals heterointerfaces,” Nat. Nanotechnol. 9(9), 676–681 (2014). [CrossRef]   [PubMed]  

83. R. Cheng, D. Li, H. Zhou, C. Wang, A. Yin, S. Jiang, Y. Liu, Y. Chen, Y. Huang, and X. Duan, “Electroluminescence and photocurrent generation from atomically sharp WSe2/MoS2 heterojunction p-n diodes,” Nano Lett. 14(10), 5590–5597 (2014). [CrossRef]   [PubMed]  

84. M. M. Furchi, A. Pospischil, F. Libisch, J. Burgdörfer, and T. Mueller, “Photovoltaic Effect in an Electrically Tunable van der Waals Heterojunction,” Nano Lett. 14(8), 4785–4791 (2014). [CrossRef]   [PubMed]  

85. K. Zhang, T. Zhang, G. Cheng, T. Li, S. Wang, W. Wei, X. Zhou, W. Yu, Y. Sun, P. Wang, D. Zhang, C. Zeng, X. Wang, W. Hu, H. J. Fan, G. Shen, X. Chen, X. Duan, K. Chang, and N. Dai, “Interlayer transition and infrared photodetection in atomically thin type-II MoTe2/MoS2 van der Waals heterostructures,” ACS Nano 10(3), 3852–3858 (2016). [CrossRef]   [PubMed]  

86. D. Xiao, G.-B. Liu, W. Feng, X. Xu, and W. Yao, “Coupled spin and valley physics in monolayers of MoS2 and other group-VI dichalcogenides,” Phys. Rev. Lett. 108(19), 196802 (2012). [CrossRef]   [PubMed]  

87. X. Xu, W. Yao, D. Xiao, and T. F. Heinz, “Spin and pseudospins in layered transition metal dichalcogenides,” Nat. Phys. 10(5), 343–350 (2014). [CrossRef]  

88. T. Cao, G. Wang, W. Han, H. Ye, C. Zhu, J. Shi, Q. Niu, P. Tan, E. Wang, B. Liu, and J. Feng, “Valley-selective circular dichroism of monolayer molybdenum disulphide,” Nat. Commun. 3, 887 (2012). [CrossRef]   [PubMed]  

89. H. Zeng, J. Dai, W. Yao, D. Xiao, and X. Cui, “Valley polarization in MoS2 monolayers by optical pumping,” Nat. Nanotechnol. 7(8), 490–493 (2012). [CrossRef]   [PubMed]  

90. H. Yuan, X. Wang, B. Lian, H. Zhang, X. Fang, B. Shen, G. Xu, Y. Xu, S. C. Zhang, H. Y. Hwang, and Y. Cui, “Generation and electric control of spin-valley-coupled circular photogalvanic current in WSe2.,” Nat. Nanotechnol. 9(10), 851–857 (2014). [CrossRef]   [PubMed]  

91. K. F. Mak, K. L. McGill, J. Park, and P. L. McEuen, “Valleytronics. the valley Hall effect in MoS₂ transistors,” Science 344(6191), 1489–1492 (2014). [CrossRef]   [PubMed]  

92. J. Lee, K. F. Mak, and J. Shan, “Electrical control of the valley Hall effect in bilayer MoS2 transistors,” Nat Nano advance online publication(2016). [CrossRef]  

93. S. Wu, J. S. Ross, G.-B. Liu, G. Aivazian, A. Jones, Z. Fei, W. Zhu, D. Xiao, W. Yao, D. Cobden, and X. Xu, “Electrical tuning of valley magnetic moment through symmetry control in bilayer MoS2,” Nat. Phys. 9(3), 149–153 (2013). [CrossRef]  

94. A. Singh, G. Moody, S. Wu, Y. Wu, N. J. Ghimire, J. Yan, D. G. Mandrus, X. Xu, and X. Li, “Coherent electronic coupling in atomically thin MoSe2,” Phys. Rev. Lett. 112(21), 216804 (2014). [CrossRef]  

95. L. Yang, N. A. Sinitsyn, W. Chen, J. Yuan, J. Zhang, J. Lou, and S. A. Crooker, “Long-lived nanosecond spin relaxation and spin coherence of electrons in monolayer MoS2 and WS2,” Nat. Phys. 11(10), 830–834 (2015). [CrossRef]  

96. C. Mai, A. Barrette, Y. Yu, Y. G. Semenov, K. W. Kim, L. Cao, and K. Gundogdu, “Many-body effects in valleytronics: direct measurement of valley lifetimes in single-layer MoS2.,” Nano Lett. 14(1), 202–206 (2014). [CrossRef]   [PubMed]  

97. W.-T. Hsu, Y.-L. Chen, C.-H. Chen, P.-S. Liu, T.-H. Hou, L.-J. Li, and W.-H. Chang, “Optically initialized robust valley-polarized holes in monolayer WSe2.,” Nat. Commun. 6, 8963 (2015). [CrossRef]   [PubMed]  

98. E. A. A. Pogna, M. Marsili, D. De Fazio, S. Dal Conte, C. Manzoni, D. Sangalli, D. Yoon, A. Lombardo, A. C. Ferrari, A. Marini, G. Cerullo, and D. Prezzi, “Photo-induced bandgap renormalization governs the ultrafast response of single-layer MoS2.,” ACS Nano 10(1), 1182–1188 (2016). [CrossRef]   [PubMed]  

99. S. Kar, Y. Su, R. R. Nair, and A. K. Sood, “Probing photoexcited carriers in a few-layer MoS2 laminate by time-resolved optical pump-terahertz probe spectroscopy,” ACS Nano 9(12), 12004–12010 (2015). [CrossRef]   [PubMed]  

100. H. Schmidt, I. Yudhistira, L. Chu, A. H. Castro Neto, B. Özyilmaz, S. Adam, and G. Eda, “Quantum transport and observation of Dyakonov-Perel spin-orbit scattering in monolayer MoS_2,” Phys. Rev. Lett. 116(4), 046803 (2016). [CrossRef]   [PubMed]  

101. L. Yang, W. Chen, K. M. McCreary, B. T. Jonker, J. Lou, and S. A. Crooker, “Spin coherence and dephasing of localized electrons in monolayer MoS₂,” Nano Lett. 15(12), 8250–8254 (2015). [CrossRef]   [PubMed]  

102. H. Wang, C. Zhang, and F. Rana, “Surface recombination limited lifetimes of photoexcited carriers in few-layer transition metal dichalcogenide MoS₂,” Nano Lett. 15(12), 8204–8210 (2015). [CrossRef]   [PubMed]  

103. C. H. Lui, A. J. Frenzel, D. V. Pilon, Y. H. Lee, X. Ling, G. M. Akselrod, J. Kong, and N. Gedik, “Trion-induced negative photoconductivity in monolayer MoS2.,” Phys. Rev. Lett. 113(16), 166801 (2014). [CrossRef]   [PubMed]  

104. Z. Nie, R. Long, L. Sun, C.-C. Huang, J. Zhang, Q. Xiong, D. W. Hewak, Z. Shen, O. V. Prezhdo, and Z.-H. Loh, “Ultrafast carrier thermalization and cooling dynamics in few-layer MoS2.,” ACS Nano 8(10), 10931–10940 (2014). [CrossRef]   [PubMed]  

105. R. Schmidt, G. Berghäuser, R. Schneider, M. Selig, P. Tonndorf, E. Malić, A. Knorr, S. Michaelis de Vasconcellos, and R. Bratschitsch, “Ultrafast coulomb-induced intervalley coupling in atomically thin WS2.,” Nano Lett. 16(5), 2945–2950 (2016). [CrossRef]   [PubMed]  

106. H. Wang, C. Zhang, and F. Rana, “Ultrafast dynamics of defect-assisted electron-hole recombination in monolayer MoS2.,” Nano Lett. 15(1), 339–345 (2015). [CrossRef]   [PubMed]  

107. J. Kim, X. Hong, C. Jin, S. F. Shi, C. Y. S. Chang, M. H. Chiu, L. J. Li, and F. Wang, “Ultrafast generation of pseudo-magnetic field for valley excitons in WSe₂ monolayers,” Science 346(6214), 1205–1208 (2014). [CrossRef]   [PubMed]  

108. L. Xuefeng, Y. Hongyi, J. Qingqing, G. Zhihan, G. Shaofeng, Q. Jun, L. Zhongfan, Z. Yanfeng, and S. Dong, “An ultrafast terahertz probe of the transient evolution of the charged and neutral phase of photo-excited electron-hole gas in a monolayer semiconductor,” 2d Mater. 3(1), 014001 (2016). [CrossRef]  

109. C. J. Docherty, P. Parkinson, H. J. Joyce, M.-H. Chiu, C.-H. Chen, M.-Y. Lee, L.-J. Li, L. M. Herz, and M. B. Johnston, “Ultrafast transient terahertz conductivity of monolayer MoS2 and WSe2 grown by chemical vapor deposition,” ACS Nano 8(11), 11147–11153 (2014). [CrossRef]   [PubMed]  

110. Q. Wang, S. Ge, X. Li, J. Qiu, Y. Ji, J. Feng, and D. Sun, “Valley carrier dynamics in monolayer molybdenum disulfide from helicity-resolved ultrafast pump-probe spectroscopy,” ACS Nano 7(12), 11087–11093 (2013). [CrossRef]   [PubMed]  

111. E. J. Sie, J. W. McIver, Y.-H. Lee, L. Fu, J. Kong, and N. Gedik, “Valley-selective optical Stark effect in monolayer WS2.,” Nat. Mater. 14(3), 290–294 (2014). [CrossRef]   [PubMed]  

112. S. Cha, J. H. Sung, S. Sim, J. Park, H. Heo, M.-H. Jo, and H. Choi, “1s-intraexcitonic dynamics in monolayer MoS2 probed by ultrafast mid-infrared spectroscopy,” Nat. Commun. 7, 10768 (2016). [CrossRef]   [PubMed]  

113. Z. He, W. Xu, Y. Zhou, X. Wang, Y. Sheng, Y. Rong, S. Guo, J. Zhang, J. M. Smith, and J. H. Warner, “Biexciton formation in bilayer tungsten disulfide,” ACS Nano 10(2), 2176–2183 (2016). [CrossRef]   [PubMed]  

114. Y. Lin, X. Ling, L. Yu, S. Huang, A. L. Hsu, Y.-H. Lee, J. Kong, M. S. Dresselhaus, and T. Palacios, “Dielectric screening of excitons and trions in single-layer MoS2.,” Nano Lett. 14(10), 5569–5576 (2014). [CrossRef]   [PubMed]  

115. K. Hao, G. Moody, F. Wu, C. K. Dass, L. Xu, C.-H. Chen, L. Sun, M.-Y. Li, L.-J. Li, A. H. MacDonald, and X. Li, “Direct measurement of exciton valley coherence in monolayer WSe2,” Nat. Phys.; advance online publication (2016). [CrossRef]  

116. D. K. Zhang, D. W. Kidd, and K. Varga, “Excited biexcitons in transition metal dichalcogenides,” Nano Lett. 15(10), 7002–7005 (2015). [CrossRef]   [PubMed]  

117. H. Shi, R. Yan, S. Bertolazzi, J. Brivio, B. Gao, A. Kis, D. Jena, H. G. Xing, and L. Huang, “Exciton dynamics in suspended monolayer and few-layer MoS2 2D crystals,” ACS Nano 7(2), 1072–1080 (2013). [CrossRef]   [PubMed]  

118. M. Palummo, M. Bernardi, and J. C. Grossman, “Exciton radiative lifetimes in two-dimensional transition metal dichalcogenides,” Nano Lett. 15(5), 2794–2800 (2015). [CrossRef]   [PubMed]  

119. W. Yao, “Excitons: Molecules in flatland,” Nat. Phys. 11(6), 448–449 (2015). [CrossRef]  

120. G. Moody, C. Kavir Dass, K. Hao, C.-H. Chen, L.-J. Li, A. Singh, K. Tran, G. Clark, X. Xu, G. Berghäuser, E. Malic, A. Knorr, and X. Li, “Intrinsic homogeneous linewidth and broadening mechanisms of excitons in monolayer transition metal dichalcogenides,” Nat. Commun. 6, 8315 (2015). [CrossRef]   [PubMed]  

121. K. J. Czech, B. J. Thompson, S. Kain, Q. Ding, M. J. Shearer, R. J. Hamers, S. Jin, and J. C. Wright, “Measurement of ultrafast excitonic dynamics of few-layer mos2 using state-selective coherent multidimensional spectroscopy,” ACS Nano 9(12), 12146–12157 (2015). [CrossRef]   [PubMed]  

122. P. Rivera, J. R. Schaibley, A. M. Jones, J. S. Ross, S. Wu, G. Aivazian, P. Klement, K. Seyler, G. Clark, N. J. Ghimire, J. Yan, D. G. Mandrus, W. Yao, and X. Xu, “Observation of long-lived interlayer excitons in monolayer MoSe2-WSe2 heterostructures,” Nat. Commun. 6, 6242 (2015). [CrossRef]   [PubMed]  

123. D. Sun, Y. Rao, G. A. Reider, G. Chen, Y. You, L. Brézin, A. R. Harutyunyan, and T. F. Heinz, “Observation of rapid exciton-exciton annihilation in monolayer molybdenum disulfide,” Nano Lett. 14(10), 5625–5629 (2014). [CrossRef]   [PubMed]  

124. J. R. Schaibley, T. Karin, H. Yu, J. S. Ross, P. Rivera, A. M. Jones, M. E. Scott, J. Yan, D. G. Mandrus, W. Yao, K.-M. Fu, and X. Xu, “Population pulsation resonances of excitons in monolayer MoSe2 with sub-1 μeV linewidths,” Phys. Rev. Lett. 114(13), 137402 (2015). [CrossRef]   [PubMed]  

125. A. Singh, G. Moody, K. Tran, M. E. Scott, V. Overbeck, G. Berghäuser, J. Schaibley, E. J. Seifert, D. Pleskot, N. M. Gabor, J. Yan, D. G. Mandrus, M. Richter, E. Malic, X. Xu, and X. Li, “Trion formation dynamics in monolayer transition metal dichalcogenides,” Phys. Rev. B 93(4), 041401 (2016). [CrossRef]  

126. P. Rivera, K. L. Seyler, H. Yu, J. R. Schaibley, J. Yan, D. G. Mandrus, W. Yao, and X. Xu, “Valley-polarized exciton dynamics in a 2D semiconductor heterostructure,” Science 351(6274), 688–691 (2016). [CrossRef]   [PubMed]  

127. A. Allain, J. Kang, K. Banerjee, and A. Kis, “Electrical contacts to two-dimensional semiconductors,” Nat. Mater. 14(12), 1195–1205 (2015). [CrossRef]   [PubMed]  

128. S. Jo, D. Costanzo, H. Berger, and A. F. Morpurgo, “Electrostatically induced superconductivity at the surface of WS₂,” Nano Lett. 15(2), 1197–1202 (2015). [CrossRef]   [PubMed]  

129. H.-J. Chuang, X. Tan, N. J. Ghimire, M. M. Perera, B. Chamlagain, M. M.-C. Cheng, J. Yan, D. Mandrus, D. Tománek, and Z. Zhou, “High mobility WSe2 p- and n-type field-effect transistors contacted by highly doped graphene for low-resistance contacts,” Nano Lett. 14(6), 3594–3601 (2014). [CrossRef]   [PubMed]  

130. H. C. P. Movva, A. Rai, S. Kang, K. Kim, B. Fallahazad, T. Taniguchi, K. Watanabe, E. Tutuc, and S. K. Banerjee, “High-mobility holes in dual-gated WSe2 field-effect transistors,” ACS Nano 9(10), 10402–10410 (2015). [CrossRef]   [PubMed]  

131. B. Radisavljevic and A. Kis, “Mobility engineering and a metal-insulator transition in monolayer MoS₂,” Nat. Mater. 12(9), 815–820 (2013). [CrossRef]   [PubMed]  

132. X. Shuigang, W. Zefei, L. Huanhuan, H. Yu, L. Gen, C. Xiaolong, H. Tianyi, Y. Weiguang, W. Yingying, L. Jiangxiazi, S. Junying, C. Yuan, H. Yuheng, Z. Fan, L. Rolf, C. Chun, and W. Ning, “Universal low-temperature Ohmic contacts for quantum transport in transition metal dichalcogenides,” 2d Mater. 3(2), 021007 (2016). [CrossRef]  

133. X. Wang, Y. Gong, G. Shi, W. L. Chow, K. Keyshar, G. Ye, R. Vajtai, J. Lou, Z. Liu, E. Ringe, B. K. Tay, and P. M. Ajayan, “Chemical vapor deposition growth of crystalline monolayer MoSe2.,” ACS Nano 8(5), 5125–5131 (2014). [CrossRef]   [PubMed]  

134. Y.-H. Chang, W. Zhang, Y. Zhu, Y. Han, J. Pu, J.-K. Chang, W.-T. Hsu, J.-K. Huang, C.-L. Hsu, M.-H. Chiu, T. Takenobu, H. Li, C.-I. Wu, W.-H. Chang, A. T. S. Wee, and L.-J. Li, “Monolayer MoSe2 Grown by chemical vapor deposition for fast photodetection,” ACS Nano 8(8), 8582–8590 (2014). [CrossRef]   [PubMed]  

135. Y. Xue, Y. Zhang, Y. Liu, H. Liu, J. Song, J. Sophia, J. Liu, Z. Xu, Q. Xu, Z. Wang, J. Zheng, Y. Liu, S. Li, and Q. Bao, “Scalable production of a few-layer MoS2/WS2 vertical heterojunction array and its application for photodetectors,” ACS Nano 10(1), 573–580 (2016). [CrossRef]   [PubMed]  

136. Y.-H. Lee, X.-Q. Zhang, W. Zhang, M.-T. Chang, C.-T. Lin, K.-D. Chang, Y.-C. Yu, J. T.-W. Wang, C.-S. Chang, L.-J. Li, and T.-W. Lin, “Synthesis of large-area MoS2 atomic layers with chemical vapor deposition,” Adv. Mater. 24(17), 2320–2325 (2012). [CrossRef]   [PubMed]  

137. C.-C. Huang, H. Medina, Y.-Z. Chen, T.-Y. Su, J.-G. Li, C.-W. Chen, Y.-T. Yen, Z. M. Wang, and Y.-L. Chueh, “Transfer-free growth of atomically thin transition metal disulfides using a solution precursor by a laser irradiation process and their application in low-power photodetectors,” Nano Lett. 16(4), 2463–2470 (2016). [CrossRef]   [PubMed]  

138. J. S. Ross, P. Klement, A. M. Jones, N. J. Ghimire, J. Yan, D. G. Mandrus, T. Taniguchi, K. Watanabe, K. Kitamura, W. Yao, D. H. Cobden, and X. Xu, “Electrically tunable excitonic light-emitting diodes based on monolayer WSe2 p-n junctions,” Nat. Nanotechnol. 9(4), 268–272 (2014). [CrossRef]   [PubMed]  

139. L. Xie and X. Cui, “Manipulating spin-polarized photocurrents in 2D transition metal dichalcogenides,” Proc. Natl. Acad. Sci. USA 113, 3746–3750 (2016). [CrossRef]   [PubMed]  

Cited By

Optica participates in Crossref's Cited-By Linking service. Citing articles from Optica Publishing Group journals and other participating publishers are listed here.

Alert me when this article is cited.


Figures (4)

Fig. 1
Fig. 1 (a). Crystal structure of TMD. (b). The honeycomb lattice structure and first Brillouin zone with high-symmetry points. (c).The optical selection rules in MoS2, the left circular polarized light/right circular polarized light can selected to excite carriers in K/K' valley respectively. (d). The absorption spectra and PL spectra of one-layer and two-layer MoS2, the spectra are displaced along the vertical axis for clarity.[From Mak.K.F. et. al, Phys Rev Lett 105.136805(2010)] (e). Schematic representation of photocurrent generation in biased TMD devices. (f). Schematic representation of photocurrent generation due to photothermalelectric effect.
Fig. 2
Fig. 2 (a). TPPC (Vc) as a function of the time delay (Δt) between two pulses show a transient photocurrent response with two distinct timescales (i) a fast time scale of ~4.3ps and(ii) a slow timescale of ~105ps. [From Wang, H. N. et al., Nature Commu 6:8831(2015).] (b). The three temporal regions of ultrafast carrier dynamics of MoS2, (I) carriers thermalization and cooling process within 500 fs. (II) electrons and holes captured by fast defects within 1-2 ps (III) electrons captured by slow defects on a 60-70 ps timescale [From Wang, H. N. et al., Nano Lett. 15, 339−345(2015).] (c).The photoresponse of the MoS2 device can be classified into 5 stages. In addition to the photoresponse due to band-band transition (stage 2 and 4), the device exhibits a slow increase in the photocurrent under illumination (stage 3) and the PPC effect (stage 5). [From Wu, Yueh-Chun et al., Scientific Reports 5:11472(2015).] (d). The trapped electrons and holes are thermally excited to defects and result in PPC effect.
Fig. 3
Fig. 3 (a). CPGE in bilayer WSe2 with excitation photon energy (1.17eV) that is below the indirect bandgap. [From Yuan, H. T. et al., Nat Nanotechnol 9, 851-857 (2014).] (b). The σ + and σ- excitations which are on-resonance with A exciton—a hole in the upper VB and electron in the CB lead to photocurrent generation in K and -K valleys, respectively. (c). CPGE current is generated due to interband transitions with excitation photon energy on-resonance with excitonic transition (1.96eV) and disappears with off-resonance excitations. [From Eginligil, M. et al., Nat Commun. 6 (2015).] (d). The valley-dependent optical selection rules and the electrons at the K and K′ valleys that possess opposite Berry curvatures (left) and a photoinduced AHE driven by a net valley polarization (middle) and an image of the Hall bar device (right). (e). The source drain bias dependence of the Hall voltage for the monolayer device. [From Mak, K. F. et al., Science 344, 1489-1492 (2014).]
Fig. 4
Fig. 4 (a). The Schematic of a PN junction in monolayer WSe2 device controlled by two local gates. [From Baugher, Britton W. H. et al., Nat Nanotechnol 9, 262-267 (2014).] (b). The Schematic of a MoS2-GQDs heterostructure phototransistor [From Chen, C. Y. et al., Scientific Reports 5:11830 (2015).] (c).The Schematic of a MoS2/WSe2 heterostructure. [From Lee, Chul-Ho et al., Nat Nanotechnol 9,676-681 (2014).] (d). The Geometry of a MoS2/Si heterojunction diode. [From Lopez-Sanchez, Oriol et al., ACS Nano 8, 3042-3048, (2014).]

Tables (1)

Tables Icon

Table 1 Comparison of Device Performance of TMD based Photodetectors

Select as filters


Select Topics Cancel
© Copyright 2024 | Optica Publishing Group. All rights reserved, including rights for text and data mining and training of artificial technologies or similar technologies.