Next Article in Journal
Nanoenergetic Materials: Preparation, Properties, and Applications
Next Article in Special Issue
Broadband Visible Nonlinear Absorption and Ultrafast Dynamics of the Ti3C2 Nanosheet
Previous Article in Journal
Vibration Sensing Systems Based on Poly(Vinylidene Fluoride) and Microwave-Assisted Synthesized ZnO Star-Like Particles with Controllable Structural and Physical Properties
Previous Article in Special Issue
Benchmark Investigation of Band-Gap Tunability of Monolayer Semiconductors under Hydrostatic Pressure with Focus-On Antimony
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Schottky Barrier Height and Image Force Lowering in Monolayer MoS2 Field Effect Transistors

School of Electrical Engineering, Tel-Aviv University, Tel Aviv 69978, Israel
*
Author to whom correspondence should be addressed.
Nanomaterials 2020, 10(12), 2346; https://doi.org/10.3390/nano10122346
Submission received: 18 October 2020 / Revised: 18 November 2020 / Accepted: 23 November 2020 / Published: 26 November 2020
(This article belongs to the Special Issue 2D Materials for Nanoelectronics)

Abstract

:
Understanding the nature of the barrier height in a two-dimensional semiconductor/metal interface is an important step for embedding layered materials in future electronic devices. We present direct measurement of the Schottky barrier height and its lowering in the transition metal dichalcogenide (TMD)/metal interface of a field effect transistor. It is found that the barrier height at the gold/ single-layer molybdenum disulfide (MoS2) interfaces decreases with increasing drain voltage, and this lowering reaches 0.5–1 V We also show that increase of the gate voltage induces additional barrier lowering.

1. Introduction

Since the introduction of the first single-layer molybdenum disulfide (MoS2) field effect transistor (FET) by Andras Kis et al. [1], in 2011, the transition metal dichalcogenides (TMDs) were proved to be beneficial as single-layer semiconductors for the post-silicon era due to their mechanical flexibility [2], high mobility and on:off ratio [3,4,5], immunity to short channel effects [6], band gap suitability to the visible-near IR spectrum [7], and abrupt switching capabilities [8]. MoS2, in particular, is considered applicable for a vast range of microelectronics [5,9], optoelectronics [10,11], sensing [12,13], spintronics [14], solar cells [15] and flexible electronics [16,17].
One of the main parameters governing the transistor characteristics, and correspondingly its potential applications, is the nature of the metal—semiconductor contact [18]. A Schottky barrier will be formed at such interfaces in the presence of a higher metal work function, as introduced by the Schottky–Mott theory [19,20]. Interface states also play a crucial role in Schottky barrier formation, especially in 2D TMD-based devices, due to the lack of dangling bonds governing the band alignment at the TMD/metal contact [21,22,23]. Many efforts are invested in reducing the Schottky barrier height or even obtaining ohmic contact of such interfaces, by either introducing a doped two-dimensional (2D) TMD layer as an intermediate layer [24] or a mixed transition layer containing van der Waals junctions [23], by using Cobalt as a tunneling barrier between the h-BN Gold electrodes [25], or by implementing Nickel electrodes [26]. Evaluation of the Schottky barrier height (SBH) in such devices can be achieved by either fitting the subthreshold regime transfer characteristics at different temperatures with the thermionic emission law [27,28], introducing radiation and measuring the resulting photocurrent [29,30], conductive atomic force microscopy (C-AFM) measurements [31], or by Kelvin probe force microscopy (KPFM) measurements [23,32]. Both the transfer characteristics fitting, the photoelectric measurement, and the KPFM methods enables device level characterization, including monitoring the terminals dependency of the barrier height.
In this work, we are using the KPFM, equipped with <25 nm radius tip, as it provides direct measurement of the band profile across the device, in contrast to the alternative methods described above; the later provides solely the SBH and not the complete band profile, and are indirect. Moreover, the KPFM feedback nullify the electrostatic force, therefore minimizes tip-induced band bending effects.
Image force barrier lowering is a dominant mechanism in the case of transistor with Schottky contacts, governed by both the drain and source bias, and the gate voltage. In the presence of this mechanism, the barrier height, governing the threshold voltage, and consequently its characteristics, will be modulated [3,29,33]. Schottky diode performance is also effected by the barrier height, as it dictates the reverse saturation current, governed by both diffusion, thermionic emission and tunneling, and hence cause ideality factor (η) variations [33]. Moreover, in the case of two-dimensional heterostructures, such as TMD/graphene, the barrier height lowering can be attributed to both graphene quantum capacitance [34,35] and the image force barrier lowering effect [36,37], effecting the on:off ratio of such FET devices. As the number of MoS2 layers decreases, the on:off ratio reduction, due to the image force, becomes more significant, as already reported [38,39]. Although the SBH in 2D TMD/metal interfaces has been widely discussed in previous reports [40,41,42,43,44], to the best of our knowledge no direct measurement of the image force barrier lowering effect in thin films has been reported. We present here a direct measurement of the Schottky barrier height and its lowering due to the image force effect, induced by both the gate voltage and the source to drain bias, in single-layer MoS2 FETs.

2. Materials and Methods

Monolayer MoS2 sample was exfoliated from an MoS2 crystal supplied by Structure Probe, Inc. (SPI) Supplies (West Chester, PA, USA) using the scotch tape technique [45], and then transferred on top of a 90 nm silicon oxide (SiO2) die. Alignment marks were patterned on the die prior exfoliation using optical lithography, and optical microscope was used to identify thin MoS2 flakes by their contrast. Contact made of 50 nm gold over 3 nm titanium were then designed by E-beam lithography and evaporated using an electron-beam evaporator. Lift-off was then performed using N-Methyl-2-pyrrolidone (NMP) at 80 °C. The device was then taped onto a chip carrier using carbon tape, wire bonded and carried into an N2 glove box for further annealing and measurements. Electrical measurements, including device characteristics, performed inside the N2 glove box, were conducted using a semiconductor parameter analyzer (B1500A, Agilent Technologies, Santa Clara, CA, USA), in addition to KPFM enabling electrostatic potential measurements in-operando. Raman spectra was collected, using a 532 nm laser of the HORIBA LabRAM HR Evolution Raman spectrometer (Kyoto, Japan). A 100× objective was used to obtain <1 μm < spot diameter with 1800 g mm−1 grating.

3. Results and Discussion

When an external field ψ is applied to a metal/semiconductor system in the −x direction (see Figure 1), the total potential energy of an electron within the semiconductor is given by:ase
E ( x ) = q 2 16 π ϵ 0 ϵ e f f x q | ψ | x
where q is the electron charge, ϵ0 is the semiconductor permittivity, ψ is the electric field within the semiconductor, and x is the distance into the semiconductor away from the interface. ϵeff in the above equation represents the effective permittivity of a monolayer semiconductor, calculated as the average of the dielectric constant of its surrounding materials. When an electron is in the vicinity of a metal electrode, an attractive force, known as the image force, will be formed due to the induced positive image charge in the metal. The first term in Equation (1) represents the work required for an electron to overcome this image force, and the second term is the work due to the electric field at the junction, located at x = 0, as described by the Schottky—Mott theory [19,20].
The peak position and the barrier lowering are extracted from Equation (1), and given by [18]:
x m = q 16 π ϵ 0 ϵ e f f | ψ |               ,               Δ ϕ = q | ψ | 4 π ϵ 0 ϵ e f f = 2 | ψ | x m
where xm is the peak position, and Δϕ is the barrier lowering. The difference between the work function of the metal and the semiconductor affinity, denoted as Bn0 at the metal side in Figure 1a, is composed of: (1) The energy difference between the conduction band and the Fermi level (FL) of the semiconductor; (2) the semiconductor band bending denoted as bi; and (3) the image force induced barrier lowering denoted as qΔϕ. The red and blue dashed lines in Figure 1a represent the image charges and the barrier height-induced electric field contribution to the potential energy of the electron as a function of its distance from the metal/semiconductor interface, respectively.
Figure 1b depicts our experimental set up, which is a monolayer MoS2 FET composed of a channel, drain and source gold electrodes, and a back-gate, in addition to the measuring KPFM tip. The device was scanned by the KPFM tip, using the dual frequency technique, to simultaneously obtain the topography and the electrostatic potential distribution along the device. An AFM image of the measured FET device, in addition to a topography cross-section is presented in Figure 1c. Figure 1d presents measured (purple) and fitted (pink) Raman spectrum, and shows a separation of 18.06 cm−1 between E 2 g 1 and A 1 g corresponding to single layer MoS2 [46]. I-V characteristics, including Id(Vd) and Id(Vg) curves, are presented in Figure S1 of the supporting information.
It should be emphasized that, by definition, the KPFM measurement follows the local vacuum level (LVL). This implies that in the absence of image force barrier lowering, both the LVL and the KPFM measurement represent the difference between the work functions of the metal and the semiconductor. By applying a voltage on one of the metal electrodes (source or drain), the electric field at the interface, and consequently the barrier height, will change; this will be measured by the KPFM as described in details below.
Figure 2a presents a series of KPFM measurements, showing the contact potential difference (CPD) along the monolayer MoS2 FET, for various drain voltages and VS = Vg = 0 V; a closer view of the CPD profile at the source/MoS2 interface is presented in Figure 2b. A depletion region of ~0.2 μm is observed at zero drain bias, in agreement with previous theoretical reports [47,48]. Figure 2c presents a series of KPFM measurements, showing the CPD distribution along the same device for various drain voltages, where VS = 0 V, and Vg = −3 V. The actual interface position was measured by KPFM to be 1.6 × 10−6 m as presented in Figure 1c. The relative peak position was extracted from this measurement and is attached in Figure 2d. The CPD peak potential, relative to the grounded source, is presented in Figure 2e.
The electric field distribution along the device, for both Vg = 0 V and Vg = −3 V, were calculated as the first derivative of the CPD profile measured by the KPFM and are presented in Figure S2 in the supporting information. The electric field at the metal/MoS2 interface, as a function of the applied drain voltage, for both gate voltages, is presented in Figure 2f. As the drain voltage increases, the CPD difference at the source/MoS2 contact (CPDsource − CPDMoS2) increases, and consequently the CPD slope from xm to the conducting channel increases. This CPD change is equivalent to modifying the slope of the dashed red line in Figure 1a. As the drain voltage increases, the electrostatic potential within the semiconductor approaches the drain potential. Since the source electrode is grounded, this potential shift within the semiconductor is added to the initial band bending at the source/MoS2 interface. As a result, the electric field at the interface increases, and the corresponding slope of the dashed red line becomes steeper. Therefore, as the image force remains constant, the barrier height becomes lower, and the peak moves towards the source interface. The peak height decrease with increasing drain voltage, is attributed to the segmentation change of the barrier height, Bn0 = qΔϕ + qϕbi + (VC − EF), at the semiconductor region, as presented above in Figure 1a. The actual band bending of the conduction band bi decreases, while the image force induced barrier lowering (qΔϕ) increases.
In order to compare the calculated barrier lowering to the measured one, we assume the following: A gold work function of 5.3 eV [49], MoS2 electron affinity of 4 eV [50] and an unintentionally n-type doping of the MoS2 [51], inducing an energy difference of 0.2 eV between the conduction band minimum and the FL of the MoS2 layer [52]. Consequently, the work function difference, causing the band bending at the gold/MoS2 interface, denoted as qΔϕ + qϕbi in Figure 1a, is 1.1 eV. The actual built-in potential, bi, is directly extracted from Figure 2a as the potential difference between the MoS2 layer and the peak height, and is ~0.7 eV. This means that at VDrain = 0 V, an initial barrier lowering of 0.4 eV at the gold/MoS2 interface will be added to the change in qΔϕ due to the applied source-drain voltage. The FL position, at Vs = Vd = 0 V, is shifted downward from EC as the gate voltage decreases, due to electrons attraction towards the semiconductor layer. A shift of 0.3 V, between the CPD of the MoS2 layer at Vg = 0 V and the CPD level at Vg = −3 V is presented in Figure 2a,c, in agreement with this FL shift. A dedicated scheme, presenting the CPD distribution at Vd = 0 V and in various gate voltages, and highlighting this CPS difference, is presented in Figure S3 of the supporting information. Hence, the initial barrier lowering at VDrain = 0 V and Vg = −3 V, is higher than at Vg = 0 V, and was estimated to be 0.6 eV.
Figure 3 presents a comparison of the barrier lowering, extracted directly from the measured CPD (dashed), and the calculated one using Equation (2) above (solid lines). The comparison is shown for two gate bias, −3 V and 0 V.
By increasing the drain voltage, the electric field increases, the slope of the dashed red line in Figure 1a steepen, the energy at the intersection point decreases, and consequently qΔϕ increases. Further increase of the electric field will result in negligible change of the barrier lowering and the peak position; this is due to the asymptotic behavior of the energy induced by the image charges force presented in Figure 1a. However, the FL position within the band gap and, consequently, the charge concentration within the depletion region are governed by the gate voltage. As the gate voltage increases, electrons are injected into the MoS2, and hence less positive charges will be introduced at the depletion region near the gold/MoS2 interface. The potential energy associated with the corresponding induced image charges (noted as the first term of Equation (1)) will then decrease. Therefore, the energy at the intersection point increases, and the barrier lowering decreases, as presented. This gate-dependent Schottky barrier height behavior was previously reported by Chen et al. [40]. The barrier height reduction of ~0.7 V at Vd = 1 V and Vg = −3 V, resulting with barrier height of ~0.6 V, is in accordance with previous reports [53,54], presenting mean barrier height of ~0.5 V in top contacted MoS2 devices.
Given an ideal Schottky diode, the reverse bias saturation current depends exponentially on the barrier height. By applying positive bias to the diode, the theoretical band bending, denoted as qΔϕ + bi in Figure 1a, will decrease. As a result, and based on the above observation presented in Figure 3, the barrier lowering induced by the image charges will decrease by approximately 0.1 V. The reverse saturation current will decrease due to the larger barrier height, and the ideality factor will be modulated accordingly. Furthermore, an increase of the gate voltage will reduce the barrier lowering effect as presented, and hence the barrier height will increase. The dark current will then decrease, resulting with increased ideality factor, as presented by Moon et al. [53].
In addition, as the thermionic emission in metal/semiconductor interfaces is exponentially proportional to the barrier height, the resulting on:off ratio in such FET devices depends on the change of the barrier height in those two states. Since the barrier lowering is more dominant at negative gate voltages, as presented in Figure 3, an on:off ratio reduction will be obtained, in agreement with previous reports [37,39,40]. Extraction of the image force barrier lowering contribution to the overall on:off ratio reduction enables further evaluation of the other involved mechanisms in 2D TMD-based FET devices. Moreover, by knowing the energy separation between the FL and the conduction band minimum of a semiconductor, this technique may be used to evaluate the exact Schottky barrier height formed in a certain device.

4. Conclusions

In summary, we have used KPFM to measure both the image force barrier lowering effect in addition to the built-in potential in thin-film MoS2 FET devices. We have also shown good agreement between the measured change in the barrier modulation and the expected lowering according to the literature. It was found that the Schottky barrier height decreases with increasing gate voltage, consistent with previous reports. Better understanding of the image force barrier lowering effect, as well as evaluation of the Schottky barrier height, in thin-film TMD-based transistors, helps with improving the contact quality of such devices.

Supplementary Materials

The following are available online at https://www.mdpi.com/2079-4991/10/12/2346/s1. Figure S1. (a) Id(Vd) characteristics. (b) Id(Vg) characteristics. Figure S2. Electric field distribution at the gold/MoS2 interface periphery of the source contact, for both VBack–gate = 0 V and VBack–gate = −3 V, respectively. The electric fields were calculated as the first derivation of the CPD distribution presented in Figure 2a,c of the manuscript, measured by the KPFM. Figure S3. CPD profile of both Vg = 0 V and Vg = −3 V at Vd = Vs = 0 V presenting separations of 0.3 eV are presented.

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.; validation Y.R.; supervision Y.R.; writing—review and editing Y.R. All authors have read and agreed to the published version of the manuscript.

Funding

This research was supported by ISF grant 537/17.

Acknowledgments

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

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Radisavljevic, B.; Radenovic, A.; Brivio, J.; Giacometti, V.; Kis, A. Single-layer MoS2 transistors. Nat. Nanotechnol. 2011, 6, 147–150. [Google Scholar] [CrossRef] [PubMed]
  2. Yoon, J.; Park, W.; Bae, G.Y.; Kim, Y.; Jang, H.S.; Hyun, Y.; Lim, S.K.; Kahng, Y.H.; Hong, W.K.; Lee, B.H.; et al. Highly flexible and transparent multilayer MoS2 transistors with graphene electrodes. Small 2013, 9, 3295–3300. [Google Scholar] [CrossRef] [PubMed]
  3. Shih, C.J.; Wang, Q.H.; Son, Y.; Jin, Z.; Blankschtein, D.; Strano, M.S. Tuning on-off current ratio and field-effect mobility in a MoS2-graphene heterostructure via schottky barrier modulation. ACS Nano 2014, 8, 5790–5798. [Google Scholar] [CrossRef] [PubMed]
  4. Wu, W.; De, D.; Chang, S.C.; Wang, Y.; Peng, H.; Bao, J.; Pei, S.S. High mobility and high on/off ratio field-effect transistors based on chemical vapor deposited single-crystal MoS2 grains. Appl. Phys. Lett. 2013, 102, 142106. [Google Scholar] [CrossRef] [Green Version]
  5. 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] [Green Version]
  6. Liu, H.; Neal, A.T.; Ye, P.D. Channel Length Scaling of MoS2 MOSFETs. ACS Nano 2012, 6, 8563–8569. [Google Scholar] [CrossRef] [Green Version]
  7. Splendiani, A.; Sun, L.; Zhang, Y.; Li, T.; Kim, J.; Chim, C.Y.; Galli, G.; Wang, F. Emerging photoluminescence in monolayer MoS2. Nano Lett. 2010, 10, 1271–1275. [Google Scholar] [CrossRef]
  8. Yoon, Y.; Ganapathi, K.; Salahuddin, S. How good can monolayer MoS2 transistors be? Nano Lett. 2011, 11, 3768–3773. [Google Scholar] [CrossRef]
  9. 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]
  10. 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]
  11. 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]
  12. He, Q.; Zeng, Z.; Yin, Z.; Li, H.; Wu, S.; Huang, X.; Zhang, H. Fabrication of flexible MoS2 thin-film transistor arrays for practical gas-sensing applications. Small 2012, 8, 2994–2999. [Google Scholar] [CrossRef] [PubMed]
  13. Liu, B.; Chen, L.; Liu, G.; Abbas, A.N.; Fathi, M.; Zhou, C. High-performance chemical sensing using Schottky-contacted chemical vapor deposition grown monolayer MoS2 transistors. ACS Nano 2014, 8, 5304–5314. [Google Scholar] [CrossRef] [PubMed]
  14. Klinovaja, J.; Loss, D. Spintronics in MoS2 monolayer quantum wires. Phys. Rev. B Condens. Matter Mater. Phys. 2013, 88, 1–6. [Google Scholar] [CrossRef] [Green Version]
  15. 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]
  16. Li, X.; Zhu, H. Two-dimensional MoS2: Properties, preparation, and applications. J. Mater. 2015, 1, 33–44. [Google Scholar] [CrossRef] [Green Version]
  17. Akinwande, D.; Petrone, N.; Hone, J. Two-dimensional flexible nanoelectronics. Nat. Commun. 2014, 5, 1–12. [Google Scholar] [CrossRef]
  18. Li, S.S. Metal–Semiconductor Contacts. In Semiconductor Physical Electronics; Springer: New York, NY, USA, 2006; pp. 284–333. [Google Scholar]
  19. Schottky, W. Zur Halbleitertheorie der Sperrschicht- und Spitzengleichrichter. Z. Phys. 1939, 113, 367–414. [Google Scholar] [CrossRef]
  20. Mott, N.F. The theory of crystal rectifiers. Proc. R. Soc. Lond. Ser. A Math. Phys. Sci. 1939, 171, 27–38. [Google Scholar]
  21. Mönch, W. Valence-band offsets and Schottky barrier heights of layered semiconductors explained by interface-induced gap states. Appl. Phys. Lett. 1998, 72, 1899–1901. [Google Scholar] [CrossRef]
  22. Fan, Y.; Zhou, Y.; Wang, X.; Tan, H.; Rong, Y.; Warner, J.H. Photoinduced Schottky Barrier Lowering in 2D Monolayer WS2 Photodetectors. Adv. Opt. Mater. 2016, 4, 1573–1581. [Google Scholar] [CrossRef]
  23. Kim, Y.; Kim, A.R.; Yang, J.H.; Chang, K.E.; Kwon, J.D.; Choi, S.Y.; Park, J.; Lee, K.E.; Kim, D.H.; Choi, S.M.; et al. Alloyed 2D Metal-Semiconductor Heterojunctions: Origin of Interface States Reduction and Schottky Barrier Lowering. Nano Lett. 2016, 16, 5928–5933. [Google Scholar] [CrossRef] [PubMed]
  24. Chuang, H.J.; Chamlagain, B.; Koehler, M.; Perera, M.M.; Yan, J.; Mandrus, D.; Tománek, D.; Zhou, Z. Low-Resistance 2D/2D Ohmic Contacts: A Universal Approach to High-Performance WSe2, MoS2, and MoSe2 Transistors. Nano Lett. 2016, 16, 1896–1902. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Cui, X.; Shih, E.M.; Jauregui, L.A.; Chae, S.H.; Kim, Y.D.; Li, B.; Seo, D.; Pistunova, K.; Yin, J.; Park, J.H.; et al. Low temperature Ohmic Contact to Monolayer MoS2 by van der Waals Bonded Co/h-BN Electrodes. Nano Lett. 2017, 17, 4781–4786. [Google Scholar] [CrossRef]
  26. Neal, A.T.; Liu, H.; Gu, J.J.; Ye, P.D. Metal contacts to MoS2: A two-dimensional semiconductor. In Proceedings of the Device Research Conference—Conference Digest, DRC, University Park, PA, USA, 18–20 June 2012; pp. 65–66. [Google Scholar]
  27. Giannazzo, F.; Fisichella, G.; Piazza, A.; Di Franco, S.; Greco, G.; Agnello, S.; Roccaforte, F. Impact of contact resistance on the electrical properties of MoS2 transistors at practical operating temperatures. Beilstein J. Nanotechnol. 2017, 8, 254–263. [Google Scholar] [CrossRef] [Green Version]
  28. Yang, H.; Heo, J.; Park, S.; Song, H.J.; Seo, D.H.; Byun, K.E.; Kim, P.; Yoo, I.K.; Chung, H.J.; Kim, K. Graphene barristor, a triode device with a gate-controlled Schottky barrier. Science 2012, 336, 1140–1143. [Google Scholar] [CrossRef] [Green Version]
  29. Rhoderick, E.H.; Williams, R.H. Metal-Semiconductor Contacts, 2nd ed.; Clarendon Press: Oxford, UK, 1988. [Google Scholar]
  30. Sze, S.M.; Crowell, C.R.; Kahng, D. Photoelectric Determination of the Image Force Dielectric Constant for Hot Electrons in Schottky Barriers. J. Appl. Phys. 1964, 35, 2534. [Google Scholar] [CrossRef]
  31. Giannazzo, F.; Schilirò, E.; Greco, G.; Roccaforte, F. Conductive atomic force microscopy of semiconducting transition metal dichalcogenides and heterostructures. Nanomaterials 2020, 10, 803. [Google Scholar] [CrossRef]
  32. Vilarinho, P.M.; Rosenwaks, Y.; Angus, K. Scanning Probe Microscopy: Characterization, Nanofabrication and Device Application of Functional Materials; Kluwer Academic Publishers: Dordrecht, The Netherlands, 2002. [Google Scholar]
  33. Nelson, J. The Physics of Solar Cells; Imperial College Press: London, UK, 2003; ISBN 1860943403. [Google Scholar]
  34. Britnell, L.; Gorbachev, R.V.; Jalil, R.; Belle, B.D.; Schedin, F.; Mishchenko, A.; Georgiou, T.; Katsnelson, M.I.; Eaves, L.; Morozov, S.V.; et al. Field-effect tunneling transistor based on vertical graphene heterostructures. Science 2012, 335, 947–950. [Google Scholar] [CrossRef] [Green Version]
  35. Ponomarenko, L.A.; Belle, B.D.; Jalil, R.; Britnell, L.; Gorbachev, R.V.; Geim, A.K.; Novoselov, K.S.; Castro Neto, A.H.; Eaves, L.; Katsnelson, M.I. Field-effect control of tunneling barrier height by exploiting graphene’s low density of states. Appl. Phys. 2013, 113, 136502. [Google Scholar] [CrossRef] [Green Version]
  36. Sata, Y.; Moriya, R.; Yamaguchi, T.; Inoue, Y.; Morikawa, S.; Yabuki, N.; Masubuchi, S.; Machida, T. Modulation of Schottky barrier height in graphene/MoS2/metal vertical heterostructure with large current ON-OFF ratio. Jpn. J. Appl. Phys. 2015, 54, 04DJ04. [Google Scholar] [CrossRef] [Green Version]
  37. Tabata, H.; Sato, Y.; Oi, K.; Kubo, O.; Katayama, M. Bias-and Gate-Tunable Gas Sensor Response Originating from Modulation in the Schottky Barrier Height of a Graphene/MoS2 van der Waals Heterojunction. ACS Appl. Mater. Interfaces 2018, 10, 38387–38393. [Google Scholar] [CrossRef] [PubMed]
  38. Yu, W.J.; Li, Z.; Zhou, H.; Chen, Y.; Wang, Y.; Huang, Y.; Duan, X. Vertically stacked multi-heterostructures of layered materials for logic transistors and complementary inverters. Nat. Mater. 2013, 12, 246–252. [Google Scholar] [CrossRef] [PubMed]
  39. Moriya, R.; Yamaguchi, T.; Inoue, Y.; Morikawa, S.; Sata, Y.; Masubuchi, S.; Machida, T. Large current modulation in exfoliated-graphene/MoS2/metal vertical heterostructures. Appl. Phys. Lett. 2014, 105, 083119. [Google Scholar] [CrossRef] [Green Version]
  40. Chen, J.R.; Odenthal, P.M.; Swartz, A.G.; Floyd, G.C.; Wen, H.; Luo, K.Y.; Kawakami, R.K. Control of Schottky barriers in single layer MoS2 transistors with ferromagnetic contacts. Nano Lett. 2013, 13, 3106–3110. [Google Scholar] [CrossRef] [PubMed]
  41. Farmanbar, M.; Brocks, G. Controlling the Schottky barrier at MoS2/metal contacts by inserting a BN monolayer. Phys. Rev. B Condens. Matter Mater. Phys. 2015, 91, 1–5. [Google Scholar] [CrossRef] [Green Version]
  42. Lee, S.; Tang, A.; Aloni, S.; Philip Wong, H.S. Statistical Study on the Schottky Barrier Reduction of Tunneling Contacts to CVD Synthesized MoS2. Nano Lett. 2016, 16, 276–281. [Google Scholar] [CrossRef]
  43. Di Bartolomeo, A.; Grillo, A.; Urban, F.; Iemmo, L.; Giubileo, F.; Luongo, G.; Amato, G.; Croin, L.; Sun, L.; Liang, S.J.; et al. Asymmetric Schottky Contacts in Bilayer MoS2 Field Effect Transistors. Adv. Funct. Mater. 2018, 28, 1–10. [Google Scholar] [CrossRef] [Green Version]
  44. Somvanshi, D.; Kallatt, S.; Venkatesh, C.; Nair, S.; Gupta, G.; Anthony, J.K.; Karmakar, D.; Majumdar, K. Nature of carrier injection in metal/2D-semiconductor interface and its implications for the limits of contact resistance. Phys. Rev. B 2017, 96, 1–12. [Google Scholar] [CrossRef] [Green Version]
  45. Novoselov, K.S.; Geim, A.K.; Morozov, S.V.; Jiang, D.; Zhang, Y.; Dubonos, S.V.; Grigorieva, I.V.; Firsov, A.A. Electric field in atomically thin carbon films. Science 2004, 306, 666–669. [Google Scholar] [CrossRef] [Green Version]
  46. 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]
  47. 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]
  48. 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]
  49. Sachtler, W.M.H.; Dorgelo, G.J.H.; Holscher, A.A. The work function of gold. Surf. Sci. 1966, 5, 221–229. [Google Scholar] [CrossRef]
  50. 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] [Green Version]
  51. 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]
  52. Dagan, R.; Vaknin, Y.; Rosenwaks, Y. Gap state distribution and Fermi level pinning in monolayer to multilayer MoS2 field effect transistors. Nanoscale 2020, 12, 8883–8889. [Google Scholar] [CrossRef]
  53. Moon, B.H.; Han, G.H.; Kim, H.; Choi, H.; Bae, J.J.; Kim, J.; Jin, Y.; Jeong, H.Y.; Joo, M.-K.; Lee, Y.H.; et al. Junction-Structure-Dependent Schottky Barrier Inhomogeneity and Device Ideality of Monolayer MoS2 Field-Effect Transistors. ACS Appl. Mater. Interfaces 2017, 9, 11240–11246. [Google Scholar] [CrossRef]
  54. John, A.P.; Thenapparambil, A.; Thalakulam, M. Strain-engineering the Schottky barrier and electrical transport on MoS2. Nanotechnology 2020, 31, 275703. [Google Scholar] [CrossRef] [Green Version]
Figure 1. (a) Schematic layout of the energy band diagram of a metal/semiconductor junction. The red and blue dashed lines represent the image charges, and the Schottky barrier height contribution to the potential energy of an electron, respectively. (b) Schematics of a field- effect transistor (FET) composed of a monolayer molybdenum disulfide (MoS2) layer as the channel, drain and source gold terminals, and a back-gate, in addition to the Kelvin probe force microscopy (KPFM) tip. (c) An atomic force microscopy (AFM) image, in addition to a topography profile along the dashed blue line, of the single layer MoS2 device. (d) Raman spectrum, in addition to the fitted peaks, of the single layer MoS2 flake, presenting separation of 18.06 (cm−1) between E 2 g 1 and A 1 g .
Figure 1. (a) Schematic layout of the energy band diagram of a metal/semiconductor junction. The red and blue dashed lines represent the image charges, and the Schottky barrier height contribution to the potential energy of an electron, respectively. (b) Schematics of a field- effect transistor (FET) composed of a monolayer molybdenum disulfide (MoS2) layer as the channel, drain and source gold terminals, and a back-gate, in addition to the Kelvin probe force microscopy (KPFM) tip. (c) An atomic force microscopy (AFM) image, in addition to a topography profile along the dashed blue line, of the single layer MoS2 device. (d) Raman spectrum, in addition to the fitted peaks, of the single layer MoS2 flake, presenting separation of 18.06 (cm−1) between E 2 g 1 and A 1 g .
Nanomaterials 10 02346 g001
Figure 2. (a) Contact potential difference (CPD) measurements of a monolayer molybdenum disulfide (MoS2) field-effect transistor (FET), at Vg = 0 V and varying drain voltages Vd = −2 V to Vd = +7 V. (b) Closer view of the CPD distribution at the gold/MoS2 interface (at the source side). The squares represent the peak point in each curve. (c) CPD measurements of the same device, at Vg = −3 V and varying drain voltages Vd = −2 V to Vd = +7 V. (df) Peak position, CPD value at the peak position, and electric field at the gold/MoS2 interface, as a function of the applied drain voltage, for both applied gate voltages, respectively.
Figure 2. (a) Contact potential difference (CPD) measurements of a monolayer molybdenum disulfide (MoS2) field-effect transistor (FET), at Vg = 0 V and varying drain voltages Vd = −2 V to Vd = +7 V. (b) Closer view of the CPD distribution at the gold/MoS2 interface (at the source side). The squares represent the peak point in each curve. (c) CPD measurements of the same device, at Vg = −3 V and varying drain voltages Vd = −2 V to Vd = +7 V. (df) Peak position, CPD value at the peak position, and electric field at the gold/MoS2 interface, as a function of the applied drain voltage, for both applied gate voltages, respectively.
Nanomaterials 10 02346 g002
Figure 3. Barrier lowering as function of the applied drain voltage, for three different gate voltages. The squares represent the barrier lowering calculated using Equation (2), and the triangles represent the measured one.
Figure 3. Barrier lowering as function of the applied drain voltage, for three different gate voltages. The squares represent the barrier lowering calculated using Equation (2), and the triangles represent the measured one.
Nanomaterials 10 02346 g003
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Vaknin, Y.; Dagan, R.; Rosenwaks, Y. Schottky Barrier Height and Image Force Lowering in Monolayer MoS2 Field Effect Transistors. Nanomaterials 2020, 10, 2346. https://doi.org/10.3390/nano10122346

AMA Style

Vaknin Y, Dagan R, Rosenwaks Y. Schottky Barrier Height and Image Force Lowering in Monolayer MoS2 Field Effect Transistors. Nanomaterials. 2020; 10(12):2346. https://doi.org/10.3390/nano10122346

Chicago/Turabian Style

Vaknin, Yonatan, Ronen Dagan, and Yossi Rosenwaks. 2020. "Schottky Barrier Height and Image Force Lowering in Monolayer MoS2 Field Effect Transistors" Nanomaterials 10, no. 12: 2346. https://doi.org/10.3390/nano10122346

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