Next Article in Journal
Ethanol-Quenching Introduced Oxygen Vacancies in Strontium Titanate Surface and the Enhanced Photocatalytic Activity
Next Article in Special Issue
Effect of Piezoresistive Behavior on Electron Emission from Individual Silicon Carbide Nanowire
Previous Article in Journal
Multiwalled Carbon Nanotube Reinforced Bio-Based Benzoxazine/Epoxy Composites with NIR-Laser Stimulated Shape Memory Effects
Previous Article in Special Issue
Probing the Optical Properties of MoS2 on SiO2/Si and Sapphire Substrates
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Pinch-Off Formation in Monolayer and Multilayers MoS2 Field-Effect Transistors

School of Electrical Engineering, Tel-Aviv University, Tel-Aviv 69978, Israel
*
Author to whom correspondence should be addressed.
Nanomaterials 2019, 9(6), 882; https://doi.org/10.3390/nano9060882
Submission received: 7 May 2019 / Revised: 30 May 2019 / Accepted: 10 June 2019 / Published: 14 June 2019

Abstract

:
The discovery of layered materials, including transition metal dichalcogenides (TMD), gives rise to a variety of novel nanoelectronic devices, including fast switching field-effect transistors (FET), assembled heterostructures, flexible electronics, etc. Molybdenum disulfide (MoS2), a transition metal dichalcogenides semiconductor, is considered an auspicious candidate for the post-silicon era due to its outstanding chemical and thermal stability. We present a Kelvin probe force microscopy (KPFM) study of a MoS2 FET device, showing direct evidence for pinch-off formation in the channel by in situ monitoring of the electrostatic potential distribution along the conducting channel of the transistor. In addition, we present a systematic comparison between a monolayer MoS2 FET and a few-layer MoS2 FET regarding gating effects, electric field distribution, depletion region, and pinch-off formation in such devices.

1. Introduction

Two-dimensional (2D) materials, such as graphene and transition metal dichalcogenides (TMDs), attract extensive interest from the research community as they are considered as possible candidates for post-silicon electronics [1]. Molybdenum disulfide (MoS2), is considered a promising candidate for various applications in microelectronics [2,3,4,5,6], optoelectronics [7,8], sensing [9,10], spintronics [11,12], and many others [13,14] due to its unique electrical properties [15,16] and thermal stability. The performance of the applications listed above is entangled with specific electrostatic behavior at the interface between the MoS2 and the other compounds within the device [17].
Depletion region formation at MoS2 homojunctions, heterojunctions, and metal junctions has been modeled and investigated by many groups [17,18,19,20]. Nipane et al. [21] modeled the electrostatics of lateral junctions in atomically thin materials using line charges representation, Sohn et al. [22] investigated the electrostatic band alignment at Au/MoS2 contacts as a function of the thickness of MoS2 exfoliated flakes, and Chiu et al. [23] determined the band alignment in single-layer MoS2/WSe2 heterojunctions using micro X-ray photoelectron spectroscopy and scanning tunneling microscopy (STM).
Kelvin probe force microscopy (KPFM) is a powerful tool for the direct measurement of the surface potential of semiconductors [24]. Several groups have used this technique to monitor the work function difference between the different number of layers of MoS2 samples [25], studying the effect of the substrate on the electrostatic properties of MoS2 layers [26], and assessing the effect of gas and molecular adsorption on chemical vapor deposition (CVD)-grown MoS2 flakes [27,28,29]. Other groups have measured the built-in potential of single-layer MoS2 heterojunctions using KPFM [30] and demonstrated the electrical properties of the contact between MoS2 and different metals [31]. These measurements resemble the use of KPFM for contact resistance evaluation [32] and contact-free mobility estimation [33] in thin-film organic transistors.
Current saturation in thin-film FET devices is attributed to both velocity saturation and pinch-off formation [6,34]—moreover, it is a major accelerator of the device’s performance, controlling the output conductance and on/off ratio of the transistor [3]. However, no experimental results on the pinch-off effect in ultrathin TMDs have been published. In this work we present a method for the direct observation and determination of the pinch-off voltage of a thin-layer TMD FET through the measurement of the electrostatic potential along the conducting channel of the transistor in operando. We observe, for the first time to our knowledge, pinch-off formation in such devices and discuss the differences between single- and few-layer MoS2 FET devices.

2. Materials and Methods

Monolayer and few-layer MoS2 samples were transferred on top of an 8 mm square dye made of a highly doped P-type silicon wafer covered by a 90 nm silicon oxide (SiO2) layer via the mechanical exfoliation with scotch tape, initially developed for graphene [35], of MoS2 crystals supplied by Structure Probe Inc. (SPI) Supplies (West Chester, PA, USA). The wafer was patterned with gold alignment marks prior to exfoliation using optical lithography, and monolayer and few-layer MoS2 flakes were identified by their contrast using an optical microscope. Contacts made of 50 nm gold on top of a 3 nm titanium were designed by E-beam lithography and evaporated using an electron-beam evaporator (VST, Israel), and lift off was conducted using N-Methyl-2-pyrrolidone (NMP) at 80 °C. The devices were then placed on a chip carrier, wire bonded, transferred into a N2 glove box and annealed at 95 °C to reduce humidity.
The electrical measurements and device characteristics, performed inside a N2 glove box, were conducted using a semiconductor parameter analyzer (B1500A, Agilent Technologies, Santa Clara, CA, USA) and atomic force microscope (AFM)-based amplitude modulation-Kelvin probe force microscopy (AM-KPFM) (Bruker, MA, USA). The measuring step size was ~20 nm, enabling in situ measurements of the electrostatic potential with lateral resolution approaching the limit of the KPFM in nitrogen atmosphere [36].

3. Results and Discussion

An AFM image of a monolayer MoS2 FET with a channel length (L) of 5.5 um and a width (W) of 0.75 um, in addition to a topography profile along the device, are shown in Figure 1a. Raman spectroscopy, presenting a separation of 18.06 cm−1 between E 2 g 1 and A 1 g corresponding to a single layer of MoS2 [37], is presented in Figure 1b. Figure 1c shows the I d ( V d ) characteristics of the device for several gate voltages ( V g ), while Figure 1d depicts the I d ( V g ) curves in linear scale for several drain voltages ( V d ) showing n-type transistor behavior, as expected from unintentionally doped exfoliated MoS2 flakes [38,39,40]. Logarithmic scale I d ( V g ) curves, for a gate voltage range of 0.5 to 8 V in which the high drain voltage curves are united into a single curve and separated as the gate voltage increases, as expected in the presence of pinch-off, are presented in the inset. Standard metal–oxide–semiconductor field-effect transistor (MOSFET) theory [41] was used to extract the field-effect mobility from the I d ( V g ) curves in the linear regime, and found to be 2.52 cm2/V∙s. Current saturation, as depicted in the inset of Figure 1c for gate voltages of 0 and 5 V, is likely to take place in FET devices, being more possible evidence of pinch-off formation near the drain electrode.
The electrostatic potential distribution along a single-layer MoS2 FET device, as a function of the biased electrodes—source, drain, and gate, measured in situ using KPFM—is presented in Figure 2. Figure 2a shows the electrostatic potential along the conducting channel for several drain voltages where the gate is grounded. It is observed that a depletion region, resulting in an increased electrostatic potential slope near the drain electrode, starts to form at a drain voltage of 3 V, while starting at V d = 6   V the separation between the different electrostatic potential curves becomes narrower and the potential drop near the drain electrode becomes steeper. This abrupt voltage drop near the drain electrode is attributed to pinch-off formation near the electrode. Figure 2b presents the electrostatic potential along the device with V g = 5   V . In this case, due to the positive gate voltage and the corresponding electrostatic field, electrons from the different electrodes are injected into the conducting channel, and a larger drain voltage (of around 6 V) is consequently required for the formation of the depletion region.
Following the Schottky–Mott theory [42,43], the barrier height in the metal–MoS2 interface is determined by the difference between the work function of the metal and the electron affinity of the MoS2. Following this model, assuming V d = V s = 0   V , the built-in potential, and consequently the depletion region, between the gold electrodes and the channel is determined by:
ψ b i = ϕ A u ( χ M o S 2 + ( E c E f ) q ) + V
where ϕ A u is the gold work function, χ M o S 2 is the MoS2 affinity, E c E f is the energy difference between the conduction band and the Fermi level of the MoS2, and V is the potential induced by the gate electrode. Assuming a gold work function of 5.4 eV [44], a MoS2 electron affinity of 4 eV [45], and an unintentionally n-type doping concentration of 1.4 × 10 12 [ 1 c m 2 ] [38] shifting the Fermi level position of the MoS2 towards the conduction band, a Schottky barrier height of 1.4 eV will be formed and a built-in potential resulting in a depletion region will be created. Increasing the gate voltage will increase the charge carrier concentration in the conducting channel, and the depletion region will become narrower.
The charge concentration induced by the gate is:
Q = C O X ( V g s V t V ( x ) )
where C O X is the structure capacitance per unit area, V t is the threshold voltage, and V ( x ) is the measured electrostatic potential minus the electrostatic potential at V g s = 0   V [46]. Following this equation, a point with no charge will be created near the drain when V ( x ) = V d r a i n = V g s V t ; this is commonly defined as the pinch-off. Any additional increase in the drain voltage will increase the pinch-off region, inducing a larger electric field in this region, as depicted in Figure 2c above. By increasing the gate voltage, a large concentration of carriers is injected into the channel, and the pinch-off voltage is increased accordingly.
Figure 2c shows the electric field distribution across the single-layer MoS2 FET device for drain voltages between 1 and 12 V, for V g s = 0   V . The electric field is calculated as the first derivative of the electrostatic potential distribution, measured by the KPFM, as presented in Figure 2a. The inset of Figure 2c presents the electric field at the source–MoS2 and MoS2–drain interfaces as a function of increasing V d s . At drain voltages lower than the pinch-off voltage, the electric field is accumulated at the source–MoS2 interface, while at some other voltages, namely the pinch-off voltage, the electric field at the source–MoS2 interface remains constant and the electric field at the MoS2–drain interface increases. A vertical line marking the drain voltage at which the electric field becomes 1 × 10 6   V / m , representing the pinch-off voltage, is also shown in Figure 2c. Given a gate voltage of 0   V , the threshold voltage can be calculated using the following formulation: V t = V p i n c h o f f 5   V .
As the electric field at the MoS2–drain interface increases, the portion of depleted channel required to screen this field increases. In order to establish a rigorous method for pinch-off positioning, we determined the pinch-off point in a manner similar to how the slope transition is extracted in logic device state determination. Following this method, the pinch-off position is defined as the intersection between the linear fit of the plateau region of the electrostatic potential along the conducting channel and the linear fit of the steepest electrostatic potential slope at the MoS2–drain interface. The results are presented in Figure 2d, while the inset of Figure 2d shows the pinch-off position as a function of the drain voltage. As expected, the pinch-off position remains constant for drain voltages lower than the pinch-off voltage and moves towards the source as the drain voltage increases beyond this voltage.
Figure 3 shows a detailed comparison between single-layer and multilayer (composed of 4–5 layers and exhibiting 20 cm2/V∙s field-effect mobility) MoS2 FET devices. Figure 3a shows the accumulated electric field at the source–MoS2 interface as a function of the drain voltage for several gate voltages for the monolayer MoS2 FET device; the vertical lines represent the pinch-off voltage as described above. The threshold voltage, extracted using the above formula as a function of the gate voltage, is shown in the inset. Figure 3b demonstrates the pinch-off widening at the MoS2–drain interface as a function of the drain voltage for several gate voltages for the monolayer device, where the zero position is the pinch-off location for V d = 0   V and the negative values represent pinch-off widening into the conducting channel away from the drain. As expected, an increased gate voltage, resulting in enhanced conducting channel, requires a higher drain voltage to achieve pinch-off.
Figure 3c,d show the accumulated electric field at the source–MoS2 interface, in addition to the corresponding threshold voltages, and demonstrate the pinch-off widening at the MoS2–drain interface for a multilayer MoS2 FET device. As shown, the corresponding pinch-off voltages in the multilayer device are smaller relative to the monolayer device. This decrease in the pinch-off voltage with an increasing number of layers may arise from the unintentional doping in the exfoliated MoS2 samples, which is known to be attributed to sulfur vacancies [39,40,47,48,49] formed mostly at interfaces, resulting in a higher carrier concentration in monolayer samples compared to multilayers [50]; this larger carrier concentration increases the threshold and pinch-off voltages.
Sulfur vacancies in addition to intentional doping of thin-film FETs will cause variations in device performance including on/off ratio, mobility, and current saturation. Observation of the pinch-off phenomenon, extraction of the pinch-off voltage, and calculation of the corresponding threshold voltage enables us to distinguish between the governing mechanisms in such devices. In addition, since the mobility in such devices is known to be field dependent [34] at drain voltages higher than the pinch-off voltage, the electric field along the device will remain constant, and hence the mobility will be independent of the drain bias. We have presented good agreement between the measured pinch-off voltage (Figure 3a) and the current saturation voltage presented in Figure 1c. Consequently, we deduce that the pinch-off phenomenon is the dominant mechanism of current saturation in the measured devices, and that the device performance is not limited by carrier velocity saturation.

4. Conclusions

In summary, a KPFM-based study was conducted to investigate the pinch-off phenomenon in thin-layer MoS2 FET devices. We presented a direct observation of pinch-off region formation in monolayer and multilayer MoS2 FETs through a detailed analysis of the electrostatic potential distribution along the devices. We showed the pinch-off dependence, in terms of both pinch-off voltage and widening of the pinch-off region, on the applied bias and derived the threshold voltage accordingly. It was shown that the pinch-off voltage decreased with the increase of the charge carrier concentration within the conducting channel, which is consistent with the increased sulfur vacancies of MoS2 surfaces compared to multilayers. Better understanding of the pinch-off phenomenon, in addition to its crucial effect on device performance, in few–layer materials is a key point for designing the next generation of TMD-based devices.

Author Contributions

Conceptualization Y.V., R.D. and Y.R., investigation Y.V. and R.D.; writing-original draft preparation Y.V.; analysis Y.V. and R.D.; validation Y.R.; supervision Y.R.; writing- review and editing Y.R.

Funding

This research was funded by ISF Grant No. 537/17.

Acknowledgments

We would like to acknowledge the Ariel Ismach group for conducting the Raman measurements.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Islam, A.E. Current Status of Reliability in Extended and beyond CMOS Devices. IEEE Trans. Device Mater. Reliab. 2016, 16, 647–666. [Google Scholar] [CrossRef]
  2. Iannaccone, G.; Bonaccorso, F.; Palacios, T.; Colombo, L.; Seabaugh, A.; Fiori, G.; Neumaier, D.; Banerjee, S.K. Electronics based on two-dimensional materials. Nat. Nanotechnol. 2014, 9, 768–779. [Google Scholar]
  3. Yoon, Y.; Ganapathi, K.; Salahuddin, S. How good can monolayer MoS2 transistors be? Nano Lett. 2011, 11, 3768–3773. [Google Scholar] [CrossRef] [PubMed]
  4. Radisavljevic, B.; Radenovic, A.; Brivio, J.; Giacometti, V.; Kis, A. Single-layer MoS2 transistors. Nat. Nanotechnol. 2011, 6, 147–150. [Google Scholar] [CrossRef] [PubMed]
  5. Liu, H.; Neal, A.T.; Ye, P.D. Channel Length Scaling of MoS2 MOSFETs. ACS Nano 2012, 6, 8563–8569. [Google Scholar] [CrossRef]
  6. 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, 1011–1017. [Google Scholar] [CrossRef] [PubMed]
  7. 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]
  8. 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]
  9. Perkins, F.K.; Friedman, A.L.; Cobas, E.; Campbell, P.M.; Jernigan, G.G.; Jonker, B.T. Chemical vapor sensing with monolayer MoS2. Nano Lett. 2013, 13, 668–673. [Google Scholar] [CrossRef]
  10. Late, D.J.; Huang, Y.K.; Liu, B.; Acharya, J.; Shirodkar, S.N.; Luo, J.; Yan, A.; Charles, D.; Waghmare, U.V.; Dravid, V.P.; et al. Sensing behavior of atomically thin-layered MoS2 transistors. ACS Nano 2013, 7, 4879–4891. [Google Scholar] [CrossRef]
  11. Xiao, D.; Liu, G.B.; Feng, W.; Xu, X.; Yao, W. Coupled spin and valley physics in monolayers of MoS2 and other group-VI dichalcogenides. Phys. Rev. Lett. 2012, 108, 196802. [Google Scholar] [CrossRef] [PubMed]
  12. Klinovaja, J.; Loss, D. Spintronics in MoS2 monolayer quantum wires. Phys. Rev. B 2013, 88, 075404. [Google Scholar] [CrossRef]
  13. Chorkendorff, I.; Bonde, J.; Jorgensen, K.P.; Horch, S.; Jaramillo, T.F.; Nielsen, J.H. Identification of Active Edge Sites for Electrochemical H2 Evolution from MoS2 Nanocatalysts. Science 2007, 317, 100–102. [Google Scholar]
  14. Tsai, M.L.; Su, S.H.; Chang, J.K.; Tsai, D.S.; Chen, C.H.; Wu, C.I.; Li, L.J.; Chen, L.J.; He, J.H. Monolayer MoS2 heterojunction solar cells. ACS Nano 2014, 8, 8317–8322. [Google Scholar] [CrossRef] [PubMed]
  15. Li, T.; Wan, B.S.; Du, G.; Zhang, B.S.; Zeng, Z.M. Electrical performance of multilayer MoS2 transistors on high-kappa Al2O3 coated Si substrates. AIP Adv. 2015, 5, 7. [Google Scholar]
  16. Li, X.; Zhu, H. Two-dimensional MoS2: Properties, preparation, and applications. J. Mater. 2015, 1, 33–44. [Google Scholar] [CrossRef]
  17. Liu, W.; Sarkar, D.; Kang, J.; Cao, W.; Banerjee, K. Impact of Contact on the Operation and Performance of Back-Gated Monolayer MoS2 Field-Effect-Transistors. ACS Nano 2015, 9, 7904–7912. [Google Scholar] [CrossRef]
  18. Gurugubelli, V.K.; Karmalkar, S. Analytical theory of the space-charge region of lateral p-n junctions in nanofilms. J. Appl. Phys. 2015, 118, 034503. [Google Scholar] [CrossRef]
  19. Chaves, F.A.; Jiménez, D. Electrostatics of two-dimensional lateral junctions. Nanotechnology 2018, 29, 275203. [Google Scholar] [CrossRef] [Green Version]
  20. Gong, C.; Colombo, L.; Wallace, R.M.; Cho, K. The Unusual Mechanism of Partial Fermi Level Pinning at Metal-MoS2 Interfaces The Unusual Mechanism of Partial Fermi Level Pinning at Metal-MoS2 Interfaces. Nano Lett. 2014, 14, 1714–1720. [Google Scholar] [CrossRef]
  21. Nipane, A.; Jayanti, S.; Borah, A.; Teherani, J.T. Electrostatics of lateral p-n junctions in atomically thin materials. J. Appl. Phys. 2017, 122, 194501. [Google Scholar] [CrossRef]
  22. Sohn, A.; Moon, H.; Kim, J.; Seo, M.; Min, K.A.; Lee, S.W.; Yoon, S.; Hong, S.; Kim, D.W. Band Alignment at Au/MoS2 Contacts: Thickness Dependence of Exfoliated Flakes. J. Phys. Chem. C 2017, 121, 22517–22522. [Google Scholar] [CrossRef]
  23. Chiu, M.-H.; Zhang, C.; Shiu, H.-W.; Chuu, C.-P.; Chen, C.-H.; Chang, C.-Y.S.; Chen, C.-H.; Chou, M.-Y.; Shih, C.-K.; Li, L.-J. Determination of band alignment in the single-layer MoS2/WSe2 heterojunction. Nat. Commun. 2015, 6, 7666. [Google Scholar] [CrossRef] [PubMed]
  24. Vilarinho, P.M.; Rosenwaks, Y.; Angus, K. Scanning Probe Microscopy: Characterization, Nanofabrication and Device Application of Functional Materials; Kluwer Academic Publishers: Norwell, MA, USA, 2002. [Google Scholar]
  25. Choi, S.; Shaolin, Z.; Yang, W. Layer-number-dependent work function of MoS2 nanoflakes. J. Korean Phys. Soc. 2014, 64, 1550–1555. [Google Scholar] [CrossRef]
  26. Robinson, B.J.; Giusca, C.E.; Gonzalez, Y.T.; DKay, N.; Kazakova, O.; Kolosov, O.V. Structural, optical and electrostatic properties of single and fewlayers MoS2: Effect of substrate. 2D Mater. 2015, 2, 15005. [Google Scholar] [CrossRef]
  27. Kim, J.H.; Lee, J.; Kim, J.H.; Hwang, C.C.; Lee, C.; Park, J.Y. Work function variation of MoS2 atomic layers grown with chemical vapor deposition: The effects of thickness and the adsorption of water/oxygen molecules. Appl. Phys. Lett. 2015, 106, 251606. [Google Scholar] [CrossRef]
  28. Lee, S.Y.; Kim, U.J.; Chung, J.; Nam, H.; Jeong, H.Y.; Han, G.H.; Kim, H.; Oh, H.M.; Lee, H.; Kim, H.; et al. Large Work Function Modulation of Monolayer MoS2 by Ambient Gases. ACS Nano 2016, 10, 6100–6107. [Google Scholar] [CrossRef] [PubMed]
  29. Varghese, J.O.; Agbo, P.; Sutherland, A.M.; Brar, V.W.; Rossman, G.R.; Gray, H.B.; Heath, J.R. The influence of water on the optical properties of single-layer molybdenum disulfide. Adv. Mater. 2015, 27, 2734–2740. [Google Scholar] [CrossRef]
  30. Chen, K.; Wan, X.; Xie, W.; Wen, J.; Kang, Z.; Zeng, X.; Chen, H.; Xu, J. Lateral Built-In Potential of Monolayer MoS2-WS2In-Plane Heterostructures by a Shortcut Growth Strategy. Adv. Mater. 2015, 27, 6431–6437. [Google Scholar] [CrossRef]
  31. Lan, F.; Tao, Q.; Li, G. Investigation of electrical properties of contact between Molybdenum disulfide and different metals. In Proceedings of the 2015 IEEE 15th International Conference on Nanotechnology (IEEE-NANO), Rome, Italy, 27–30 July 2015; pp. 617–620. [Google Scholar]
  32. Chiarella, F.; Barra, M.; Carella, A.; Parlato, L.; Sarnelli, E.; Cassinese, A. Contact-resistance effects in PDI8-CN2 n-type thin-film transistors investigated by Kelvin-probe potentiometry. Org. Electron. 2016, 28, 299–305. [Google Scholar] [CrossRef]
  33. Chianese, F.; Chiarella, F.; Barra, M.; Carella, A.; Cassinese, A. Scanning Kelvin Probe Microscopy investigation of the contact resistances and charge mobility in n-type PDIF-CN2 thin-film transistors. Org. Electron. 2018, 52, 206–212. [Google Scholar] [CrossRef]
  34. He, G.; Ghosh, K.; Singisetti, U.; Ramamoorthy, H.; Somphonsane, R.; Bohra, G.; Matsunaga, M.; Higuchi, A.; Aoki, N.; Najmaei, S.; et al. Conduction Mechanisms in CVD-Grown Monolayer MoS2 Transistors: From Variable-Range Hopping to Velocity Saturation. Nano Lett. 2015, 15, 5052–5058. [Google Scholar] [CrossRef] [PubMed]
  35. Novoselov, K.S.; Geim, A.K.; Morozov, S.V.; Jiang, D.; Zhang, Y.; Dubonos, S.V.; Grigorieva, I.V.; Firsov, A.A. Electric field effect in atomically thin carbon films. Science 2004, 306, 666–669. [Google Scholar] [CrossRef] [PubMed]
  36. Strassburg, E.; Boag, A.; Rosenwaks, Y. Reconstruction of electrostatic force microscopy images. Rev. Sci. Instrum. 2005, 76, 1–5. [Google Scholar] [CrossRef]
  37. Li, H.; Zhang, Q.; Yap, C.C.R.; Tay, B.K.; Edwin, T.H.T.; Olivier, A.; Baillargeat, D. From bulk to monolayer MoS2: Evolution of Raman scattering. Adv. Funct. Mater. 2012, 22, 1385–1390. [Google Scholar] [CrossRef]
  38. Dagan, R.; Vaknin, Y.; Henning, A.; Shang, J.Y.; Lauhon, L.J.; Rosenwaks, Y. Two-dimensional charge carrier distribution in MoS2 monolayer and multilayers. Appl. Phys. Lett. 2019, 114, 101602. [Google Scholar] [CrossRef]
  39. Suh, J.; Park, T.E.; Lin, D.Y.; Fu, D.; Park, J.; Jung, H.J.; Chen, Y.; Ko, C.; Jang, C.; Sun, Y.; et al. Doping against the native propensity of MoS2: Degenerate hole doping by cation substitution. Nano Lett. 2014, 14, 6976–6982. [Google Scholar] [CrossRef]
  40. Liu, D.; Guo, Y.; Fang, L.; Robertson, J. Sulfur vacancies in monolayer MoS2 and its electrical contacts. Appl. Phys. Lett. 2013, 103, 183113. [Google Scholar] [CrossRef]
  41. Sze, S. Semiconductor Devices: Physics and Technology; John Wiley & Sons: Hoboken, NJ, USA, 2008. [Google Scholar]
  42. Schottky, W. Zur Halbleitertheorie der Sperrschicht- und Spitzengleichrichter. Z. Phys. 1939, 113, 367–414. [Google Scholar] [CrossRef]
  43. Mott, N.F. The theory of crystal rectifiers. Proc. R. Soc. Lond. Ser. A. Math. Phys. Sci. 1939, 171, 27–38. [Google Scholar] [CrossRef] [Green Version]
  44. Skriver, H.L.; Rosengaard, N.M. Surface energy and work function of elemental metals. Phys. Rev. B 1992, 46, 7157–7168. [Google Scholar] [CrossRef] [Green Version]
  45. Walia, S.; Balendhran, S.; Wang, Y.; Ab Kadir, R.; Sabirin Zoolfakar, A.; Atkin, P.; Zhen Ou, J.; 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]
  46. Tal, O.; Rosenwaks, Y.; Roichman, Y.; Preezant, Y.; Tessler, N.; Chan, C.K.; Kahn, A. Threshold voltage as a measure of molecular level shift in organic thin-film transistors. Appl. Phys. Lett. 2006, 88, 043509. [Google Scholar] [CrossRef]
  47. Le, D.; Rawal, T.B.; Rahman, T.S. Single-layer MoS2 with sulfur vacancies: Structure and catalytic application. J. Phys. Chem. C 2014, 118, 5346–5351. [Google Scholar] [CrossRef]
  48. Noh, J.-Y.; Kim, H.; Kim, Y.-S. Stability and electronic structures of native defects in single-layer MoS2. Phys. Rev. B 2014, 89, 205417. [Google Scholar] [CrossRef]
  49. Zhou, W.; Zou, X.; Najmaei, S.; Liu, Z.; Shi, Y.; Kong, J.; Lou, J.; Ajayan, P.M.; Yakobson, B.I.; Idrobo, J.C. Intrinsic structural defects in monolayer molybdenum disulfide. Nano Lett. 2013, 13, 2615–2622. [Google Scholar] [CrossRef]
  50. Siao, M.D.; Shen, W.C.; Chen, R.S.; Chang, Z.W.; Shih, M.C.; Chiu, Y.P.; Cheng, C.M. Two-dimensional electronic transport and surface electron accumulation in MoS2. Nat. Commun. 2018, 9, 1442. [Google Scholar] [CrossRef]
Figure 1. (a) An atomic force microscope (AFM) image, in addition to a topography profile along the dashed blue line, of the single-layer molybdenum disulfide (MoS2) device. (b) Raman spectrum of the single-layer MoS2 flake, presenting a separation of 18.06 cm−1 between E 2 g 1 and A 1 g . (c) I d ( V d ) characteristics of the single-layer MoS2 field-effect transistor (FET) device for several gate voltages, in addition to the magnified image in the inset, presenting the I d ( V d ) curves for V g = 0   V and V g = 5   V in blue and red, respectively. (d) I d ( V g ) characteristics of the FET device for several drain voltages, in both linear scale and logarithmic scale in the inset.
Figure 1. (a) An atomic force microscope (AFM) image, in addition to a topography profile along the dashed blue line, of the single-layer molybdenum disulfide (MoS2) device. (b) Raman spectrum of the single-layer MoS2 flake, presenting a separation of 18.06 cm−1 between E 2 g 1 and A 1 g . (c) I d ( V d ) characteristics of the single-layer MoS2 field-effect transistor (FET) device for several gate voltages, in addition to the magnified image in the inset, presenting the I d ( V d ) curves for V g = 0   V and V g = 5   V in blue and red, respectively. (d) I d ( V g ) characteristics of the FET device for several drain voltages, in both linear scale and logarithmic scale in the inset.
Nanomaterials 09 00882 g001
Figure 2. (a) Electrostatic potential measurements at varying drain voltages ( V d = 1   V to V d = 12   V with 1   V steps), while gate and source electrodes are grounded. (b) Electrostatic potential measurements at varying drain voltages ( V d = 1   V to V d = 7   V with 1   V steps), while V g = 5   V and the source electrode is grounded. (c) Electric field distribution along the transistor corresponding to the bias conditions in (a). The inset shows the electric field at the Au/MoS2 interface near the source and drain electrodes as a function of the drain voltage at V g = 0   V . (d) Pinch-off position extraction at V g = 0   V for representative drain voltages. In the caption, the pinch-off evaluation as a function of the drain voltage is given.
Figure 2. (a) Electrostatic potential measurements at varying drain voltages ( V d = 1   V to V d = 12   V with 1   V steps), while gate and source electrodes are grounded. (b) Electrostatic potential measurements at varying drain voltages ( V d = 1   V to V d = 7   V with 1   V steps), while V g = 5   V and the source electrode is grounded. (c) Electric field distribution along the transistor corresponding to the bias conditions in (a). The inset shows the electric field at the Au/MoS2 interface near the source and drain electrodes as a function of the drain voltage at V g = 0   V . (d) Pinch-off position extraction at V g = 0   V for representative drain voltages. In the caption, the pinch-off evaluation as a function of the drain voltage is given.
Nanomaterials 09 00882 g002
Figure 3. (a) The electric field at the Au–MoS2 interface near the source electrode as a function of the drain voltage for varying gate voltages; vertical lines representing the pinch-off voltages for the different bias conditions are also presented. The inset shows the threshold voltage extracted for the different gate voltages in a single-layer MoS2 transistor. (b) Pinch-off position as a function of the drain voltage extracted for varying gate voltages in a single-layer MoS2 transistor. (c,d) are similar to (a,b), respectively, for a few-layer MoS2 FET.
Figure 3. (a) The electric field at the Au–MoS2 interface near the source electrode as a function of the drain voltage for varying gate voltages; vertical lines representing the pinch-off voltages for the different bias conditions are also presented. The inset shows the threshold voltage extracted for the different gate voltages in a single-layer MoS2 transistor. (b) Pinch-off position as a function of the drain voltage extracted for varying gate voltages in a single-layer MoS2 transistor. (c,d) are similar to (a,b), respectively, for a few-layer MoS2 FET.
Nanomaterials 09 00882 g003

Share and Cite

MDPI and ACS Style

Vaknin, Y.; Dagan, R.; Rosenwaks, Y. Pinch-Off Formation in Monolayer and Multilayers MoS2 Field-Effect Transistors. Nanomaterials 2019, 9, 882. https://doi.org/10.3390/nano9060882

AMA Style

Vaknin Y, Dagan R, Rosenwaks Y. Pinch-Off Formation in Monolayer and Multilayers MoS2 Field-Effect Transistors. Nanomaterials. 2019; 9(6):882. https://doi.org/10.3390/nano9060882

Chicago/Turabian Style

Vaknin, Yonatan, Ronen Dagan, and Yossi Rosenwaks. 2019. "Pinch-Off Formation in Monolayer and Multilayers MoS2 Field-Effect Transistors" Nanomaterials 9, no. 6: 882. https://doi.org/10.3390/nano9060882

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