Next Article in Journal
Multi-Response Optimization of Electrical Discharge Drilling Process of SS304 for Energy Efficiency, Product Quality, and Productivity
Previous Article in Journal
Improving Sodium Alginate Films Properties by Phenolic Acid Addition
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Letter

Effects of Charge Trapping at the MoS2–SiO2 Interface on the Stability of Subthreshold Swing of MoS2 Field Effect Transistors

1
High-Frequency High-Voltage Device and Integrated Circuits R&D Center, Institute of Microelectronics, Chinese Academy of Sciences, Beijing 100029, China
2
University of Chinese Academy of Sciences, Beijing 100049, China
*
Author to whom correspondence should be addressed.
Materials 2020, 13(13), 2896; https://doi.org/10.3390/ma13132896
Submission received: 30 May 2020 / Revised: 19 June 2020 / Accepted: 22 June 2020 / Published: 28 June 2020
(This article belongs to the Section Thin Films and Interfaces)

Abstract

:
The stability of the subthreshold swing (SS) is quite important for switch and memory applications in logic circuits. The SS in our MoS2 field effect transistor (FET) is enlarged when the gate voltage sweep range expands towards the negative direction. This is quite different from other reported MoS2 FETs whose SS is almost constant while varying gate voltage sweep range. This anomalous SS enlargement can be attributed to interface states at the MoS2–SiO2 interface. Moreover, a deviation of SS from its linear relationship with temperature is found. We relate this deviation to two main reasons, the energetic distribution of interface states and Fermi level shift originated from the thermal activation. Our study may be helpful for the future modification of the MoS2 FET that is applied in the low power consumption devices and circuits.

1. Introduction

Recently, transition metal dichalcogenides (TMDCs) are leading the trend of studying various two-dimensional (2D) materials. In particular, MoS2, which is a representative of the TMDCs, presents an indirect bandgap of 1.29 eV for the bulk form and transforms to direct band semiconductor (bandgap of 1.8 eV) as the number of layers decreases to one [1,2] Additionally, for a top gate MoS2 FET, the on/off ratio is up to 109 and the subthreshold swing of is as low as 65 mV/decade, which is comparable to that of its opponent SOI MOSFET, compared to typical values of 106 and close to 80 mV/decade of conventional Si CMOS technology, making MoS2 quite suitable for use in switches, logic circuits, amplifiers, and low power dissipation circumstances [3,4,5,6,7].
However, as a 2D material, because of the extremely high surface-to-volume ratio, the properties of MoS2 FETs, such as threshold voltage, hysteresis, on/off ratio, and SS, can be easily affected by interface states at the interface between the MoS2 channel and the gate dielectric [5,8]. During the sweep of gate voltage, the interface states are electrically equivalent to an additional capacitance which is first in parallel to the semiconductor capacitance, and their result is then series to the oxide capacitance by exchanging electrons with the MoS2 channel, thus sabotaging the stability of MoS2 FETs. Among the device performance stability issues caused by the interface states, the enlargement of SS will slow down the switch speed and hugely increase the static power consumption of FETs, which will hinder the use of MoS2 FETs in low-power applications [9,10].
SS is normally considered to have a linear relationship with temperature. This might be valid for top gate FETs whose gate dielectric capacitance is large enough so that the influence of interface states capacitance (Cit) could be ignored. However, for research concerns, a back gate metal-oxide-semiconductor (MOS) structure based on SiO2 dielectric is usually adopted to study the basic properties of MoS2 FETs whose SS is typically more than 1000 mV/dec [11]. Note that the SS can be hugely decreased by replacing SiO2 with a high-k dielectric [12].
Thus, for the first time, an obvious deviation of SS from its linearity with temperature is observed for the backgated MoS2 FET with 300 nm SiO2 dielectric. We explain this anomalous phenomenon with the energetic distribution of interface states density and the temperature induced Fermi level shift. According to these results, SS stability of the devices critically relies on a rational design of gate dielectric.

2. Experiments

In our study, with the scotch tape method [13,14], the device was prepared by mechanically exfoliating MoS2 thin films onto 300 nm SiO2 grown on low resistive Si substrate. Drain and source electrode pads of Ti/Au (10/100 nm) were formed through e-beam lithography, e-beam evaporation and lift-off processes in sequence. Figure 1a,b show the schematic view and colored SEM (Tescan, Brno, Czechoslovakia) image of the fabricated MoS2 FET. The channel length and width of the device were about 3 µm and 5 µm, respectively. The thickness of the MoS2 film was about 4 nm measured using a Park Systems’ atomic force microscope (AFM, Park Systems, Hyderabad, India) as shown in Figure 1c,d. The MoS2 FET was placed in a shielded probe station under vacuum condition (less than 10−5 mbar) in order to minimize the influence of water and oxide molecules in ambient air. All electrical transport characteristics were measured using an Agilent B1500 semiconductor parameter analyzer (Agilent, Santa Clara, CA, US) under dark conditions. The transfer characteristics were obtained by sweeping VBG from a negative gate voltage VBG, min to a positive gate voltage VBG, max (forward sweep), and then back to VBG, min (backward sweep) again at a constant VDS = 1 V at room temperature. Wherein, VBG, max was fixed at 50 V and VBG, min was varied from −50 V to −80 V in steps of −10 V for each measurement, respectively.

3. Results and Discussions

Figure 2a represents the transfer curves of the MoS2 FET measured at 310 K with varying VBG, mins. With VBG, min varied from −50 V to −80 V, the threshold voltage (VTH) of the device negatively shifted for both forward sweep and backward sweep. For forward sweeps, the threshold voltage of the device shifted from −30 to −43 V. For backward sweeps, the threshold voltage of the device shifted from −23 to −35 V. Meanwhile, the hysteresis of the device increased from 5 to 9 V. For MoS2 FETs, threshold voltage shift and hysteresis are mainly caused by the interface states acting as charge traps at the MoS2–SiO2 interface. These interface states are the result of dangling bonds formed at the surface of SiO2 during the growth process of SiO2. By exchanging electrons with the MoS2 channel, the interface states can be viewed as a “impurity band” in the MoS2 bandgap [15,16]. At 310 K, as the gate voltage sweeps from VBG, min to 50 V, the conduction band bends downward relative to the Fermi level. The positions of the Fermi level corresponding to VBG, mins of −50 to −80 V are shown, respectively, in the inset of Figure 2a. As the gate voltage increases, a part of energy levels of the interface states will be swept over by the Fermi level. This will cause the charge traps to capture/release electrons. When the energy levels of the interface states are below the Fermi level, it indicates the energy level are occupied by electrons as shown in Figure 2b. When the energy levels of the interface states are above the Fermi level, electrons that used to be trapped in these interface states tend to get de-trapped, as shown in Figure 2c. When a negative gate bias is applied at the beginning of forward sweep, a portion of electrons are released to the MoS2 channel, the empty traps become positively charged. This will weaken the depletion of the MoS2 channel, because the positive charges will partially “screen” the negative gate voltage. The threshold voltage will then negatively shift. For a more negative VBG, min, more positive charges will be present. Then, it will require a more negative gate bias to depletion the MoS2 channel. Therefore, the threshold voltage will keep on shifting to the negative direction. As for the hysteresis, because we used a constant gate voltage sweeping rate of 2 V/s, the accumulation time interval in one gate voltage sweep cycle is 50 s. The depletion time interval from VBG, min to 0 V in the forward sweep is shorter than the accumulation time interval, even for the VBG, min of −80 V condition. Together with the fact that the de-trapping process is much lower than electron movements in the MoS2 channel [8], the refilled traps during the accumulation process will partly screen the negative gate voltage and cause a positive shift of VTH. Thus, the hysteresis occurs.
Like hysteresis and threshold voltage shift, SS shift can also be attributed to the interface states. However, the SS shift can hardly be observed without gate bias stressing [17], because the density of states (DOS) distribution of interface states is seldom considered. In this study, DOS distribution of the interface states is assumed to coincide with Gaussian distribution and its maximum value is located at a few tens of meV below the conduction band, as shown in Figure 2 [15,18,19]. For both forward sweeps and backward sweeps, as the gate voltage sweeping range expands, the SS of the device is enlarged because more traps are depleted or in other words “activated” when larger negative gate bias is used. SS increases from 2188 to 2898 mV/dec for forward sweeps. SS increases from 648 to 1550 mV/dec for backward sweeps. For forward sweeps, the interface states density (Dit) is extracted to be 2.64 × 1012 eV−1cm−2 for VBG, min of −50 V, and Dit is extracted to be 3.49 × 1012 eV−1cm−2 for VBG, min of −80 V, by using the following equations [20]:
S S = ln 10 × k T / q × ( 1 + C i t / C O X )
D i t = C i t / q 2
where k is the Boltzmann’s constant, T is the measurement temperature, q is the electron charge, and Cox is the gate dielectric capacitance density which is 11.5 nF/cm2 (Cox = ɛ0ɛr/d; ɛr = 3.9; d = 300 nm) for the MoS2 FET.
To further study the mechanism of the SS shift, we conducted transfer characteristics measurements at a series of temperatures. Figure 3a,b show the transfer curves for VBG, mins of −50 V for forward and backward sweeps, respectively, in a temperature range from 10 to 310 K. The VTH shifts negatively as temperature increases. The transfer curves for VBG, mins from −60 to −80 V in steps of −10 V with constant VDS = 1 V were also measured. The relationship between VTH, temperature, and VBG, min is extracted and illustrated in Figure 3c,d for forward and backward sweeps, respectively. For more negative VBG, mins, the VTH shift enlarges more abruptly compared to less negative VBG, mins when the temperature is above 100 K, which means the “screening” mentioned above becomes stronger above 100 K. No gate bias stressing is applied in this experiment, therefore the electron exchange between the MoS2 channel and interface states is incomplete. For more complete delegation by larger gate bias, the “effective” Dit is higher. As the unoccupied traps are positively charged, the electrons in the MoS2 channel will suffer from Coulomb scattering. We have σ D S n α , wherein σ D S the conductance of the MoS2, 1 α 2 is the coefficient that reflects the screening of Coulomb scattering. If the Coulomb scattering is fully screened, then α = 1 . For bare Coulomb scattering, α = 2 . As I D S σ D S , V B G V T H n , the coefficient α is the slope of ln I D S versus ln ( V B G V T H ) [8]. By calculating the data from Figure 3a,b, we obtain a minimum α of 1.56 for VBG, min of −80 V and a maximum α of 1.72 for VBG, min of −50 V at 310 K. The fact that more negative VBG, min corresponds to weaker Coulomb scattering indicates the extra electrons depleted to the MoS2 channel by the larger VBG, m (more negative) is more effective on screening the Coulomb scattering caused by the unoccupied traps than causing more severe scattering. This also suggests large amount of interface states intrinsically exist at the MoS2/SiO2 interface.
Figure 4a,b show the relationship between SS and temperature extracted from Figure 3a,b for forward sweep and backward sweep, respectively. Normally, SS should have a linear relationship with temperature. However, for forward sweeps, for VBG, min of −50 V, the SS versus T curve increased from 10 K and reached a maximum mV/dec at 50 K, then dropped to a minimum at 100 K, and began to increase linearly above 100 K. The SS versus temperature curve presents an abnormal bulge in the temperature range of 10 to 100 K. For VBG, min of −60 V, the bulge expands to 250 K and SS begins to increase linearly from there on. As VBG, min further decreases to −70 V, the bulge expands at least to 310 K. For VBG, min of −80 V, the range of the bulge is similar to that for VBG, min of −70 V. Although for VBG, mins of −70 and −80 V, the SS does not show a linear increase in the temperature range of this experiment; we believe the SS will still begin to increase linearly at a higher temperature. For backward sweeps, the range of the bulge is almost the same for each VBG, min. The SS curves all begin to increase linearly at around 250 K. Moreover, the scale of the bulge is even larger for small VBG, mins compared to forward sweeps. For this abnormal deviation of SS from linearity with temperature, the energetic distribution of interface states density and the temperature induced Fermi level shift mechanisms are adopted to explain it.
Due to the trapping and releasing processes of electrons in the traps, the interface states can be viewed as a capacitor. The capacitance Cit is determined by the density of the interface states Dit. Thus, Dit is the key to study this abnormal relationship between SS and temperature. As we assume Dit has Gaussian distribution as shown Figure 2, it can be written as
D i t ( E ) = D i t ,   m a x ( E ) × f ( E )
and f (E) in Equation (3) can be written as
f ( E ) = exp [ 4 × log 2 × ( E F ( T ) E D FWHM ) 2 ]
where in Dit, max (E) is the maximum value the Dit can reach during the shift of Fermi level as temperature increases, EF (T) is the location of Fermi level as a function of temperature, and f (E) is the energetic distribution function of the interface states [19,21]. f (E) represents the possibility that a certain energy level is occupied. FWHM is the full width at half maximum of Dit. On the other hand, because of the activation of bulk charge traps (including bulk MoS2 trap charges, fixed oxide charges, and oxide trap charges inside a thick SiO2 insulator) and interface states [22], the location of the Fermi level can be modulated by temperature and can be expressed as
E F = ( E C / E D ) / 2 + ( k T / 2 ) ln ( N D / 2 N C )
E F = E C + k T ln ( N D / N C )
corresponding to low temperature weak ionization region (Equation (5)) and high temperature strong ionization region (Equation (6)), wherein EC, NC, and ND are the bottom of the conduction band, effective density of states of conduction band, and donor concentration of the interface states for MoS2, respectively. As shown in Figure 4c, the position of the Fermi level corresponding to subthreshold situation (EF, sub) shifts downward relative to the energy levels of the interface states when temperature increases from 10 to 310 K. Thus, the Dit corresponding to EF, sub will change with temperature.
Due to the Gaussian distribution, Dit has a maximum value Dit, max. For a certain temperature, the more the Fermi level approaches Dit, max, the larger the SS is and vice versa. When substituting above Equations (2)–(6) back into Equation (1), the expression of SS can be written as
S S = q ln 10 × k T × D i t ,   m a x ( E ) × exp [ 4 × log 2 × ( E F ( T ) E D FWHM ) 2 ] / C O X + k T / q × ln 10
The deduced expression for SS has two terms. The first term is a nonlinear term which corresponds to the depletion of interface states by the gate voltage, and the second term is a linear term which is only affected by temperature. Taking the situation for VBG, min of −50 V as an example, for forward sweep, from 10 to 50 K, Dit corresponding to EF, sub increases rapidly. Thus, the nonlinear term of Equation (7) dominates and the SS curve begins to increase and deviate from linearity. After 50 K, as EF, sub passed Dit, max, the nonlinear term begins to decrease. If the increasing of the linear term and the decreasing the nonlinear term reach an equilibrium, the SS will saturate. When temperature continues to increase, the drop down of Dit will experience an acceleration. The nonlinear term will dominate again during this fast drop and SS will drop with it. When the decrease of Dit slows down at a position that is far from Dit, max, another equilibrium with the linear term will be accomplished. Thus, SS will reach a minimum. As the temperature further increases, the linear term will dominate and the nonlinear term will be negligible. The SS will increase linearly along with the linear term. When EF, sub shifts downward as temperature increases, because the effective Dit is actually higher compared to the situation for VBG, min of −50 V, the SS will be generally larger, as can be seen in Figure 4a,b and the shift of EF, sub will be slowed down in the temperature dimension because more traps have to be thermally activated. Then the maximum or minimum of SS will be achieved at higher temperatures. The above discussion also explained why the deviation from linearity is hardly observed for small gate voltage sweep ranges (e.g., from −30 V to 30 V) in which case the effect of the nonlinear term is too small and is covered up by the linear term. For backward sweep, because the traps refilled when the gate voltage is positive have not fully de-trapped yet. Then, effective Dit becomes smaller compared to forward sweep. The SS for backward sweep is thus generally smaller than that for forward sweep. This result coincides with the fact that the VTH shift for backward sweep is generally smaller than that for forward sweep, which also originates from the more incomplete depletion of interface states for backward sweep. From the above discussion, we can infer that the FWHM of Dit is important for whether the deviation of SS from linearity can be observed. If the FWHM is small enough, the distribution of Dit can be viewed as a step function. The slope of SS will change abruptly to another value at a certain temperature as is observed by Park et al. [22,23]. However, if the FWHM is large, the situation will become what we observed for our MoS2 FET.

4. Conclusions

In conclusion, we have investigated the SS instability with temperature of back-gated multilayer FET device induced by interface states in the temperature range of 10–310 K. We found that the relationship between SS and temperature will deviate from linearity if the energetic distribution of the interface states is abroad enough. Because of the Fermi level shift with temperature induced by the terminal activation of bulk charge traps and interface states, and broad interface states density distribution, the capacitor effect of the interface states can greatly influence the SS instability with temperature. Moreover, we can potentially predict the relationship between SS and temperature for MoS2 FETs with other interface states situations. Our study is helpful for understanding interface properties of MoS2 FETs and has important implications for gate dielectric material modification for low power applications based on MoS2.

Author Contributions

Conceptualization, X.H., Y.Y. and Z.J.; methodology, X.H., J.S. and D.Z.; software, X.H. and Y.Y.; validation, X.H. and Z.J.; formal analysis, X.H., Y.Y. and S.P.; investigation, X.H.; resources, Z.J.; data curation, X.H.; writing—original draft preparation, X.H.; writing—review and editing, X.H. and Z.J.; visualization, X.H. and Y.Y.; supervision, Z.J.; project administration, Z.J.; funding acquisition, Z.J. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Acknowledgments

This work was subsidized by the National Natural Science Foundation of China (No. 61704189), the Common Information System Equipment Pre-Research Special Technology Project (31513020404-2), Youth Innovation Promotion Association of Chinese Academy of Sciences, and the Opening Project of Key Laboratory of Microelectronic Devices & Integrated Technology, Institute of Microelectronics, Chinese Academy of Sciences. And the APC was funded by the Opening Project of Key Laboratory of Microelectronic Devices & Integrated Technology, Institute of Microelectronics, Chinese Academy of Sciences.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Mak, K.F.; Lee, C.; Hone, J.; Shan, J.; Heinz, T.F. Atomically thin MoS2: A new direct-gap semiconductor. Phys. Rev. Lett. 2010, 105, 136805. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Parthé, E.; Gmelin, L. Gmelin Handbook of Inorganic and Organometallic Chemistry: TYPIX.Standardized Data and Crystal Chemical Characterization of Inorganic Structure Types; Springer: Berlin/Heidelberg, Germany, 1993; Volume 1. [Google Scholar]
  3. Chang, H.Y.; Yogeesh, M.N.; Ghosh, R.; Rai, A.; Sanne, A.; Yang, S.; Lu, N.; Banerjee, S.K.; Akinwande, D. Large-area monolayer MoS2 for flexible low-power RF nanoelectronics in the GHz regime. Adv. Mater. 2016, 28, 1818–1823. [Google Scholar] [CrossRef] [PubMed]
  4. Desai, S.B.; Madhvapathy, S.R.; Sachid, A.B.; Llinas, J.P.; Wang, Q.; Ahn, G.H.; Pitner, G.; Kim, M.J.; Bokor, J.; Hu, C.; et al. MoS2 transistors with 1-nanometer gate lengths. Science 2016, 354, 99–102. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  5. Illarionov, Y.Y.; Smithe KK, H.; Waltl, M.; Knobloch, T.; Pop, E.; Grasser, T. Improved hysteresis and reliability of MoS2 transistors with high-quality CVD growth and Al2O3 encapsulation. IEEE Electron. Device Lett. 2017, 38, 1763–1766. [Google Scholar] [CrossRef]
  6. Xue, F.; Chen, L.; Wang, L.; Pang, Y.; Chen, J.; Zhang, C.; Wang, Z.L. MoS2 Tribotronic transistor for smart tactile switch. Adv. Funct. Mater. 2016, 26, 2104–2149. [Google Scholar] [CrossRef]
  7. Yu, L.; El-Damak, D.; Radhakrishna, U.; Ling, X.; Zubair, A.; Lin, Y.; Zhang, Y.; Chuang, M.H.; Lee, Y.H.; Antoniadis, D.; et al. Design, modeling, and fabrication of chemical vapor deposition grown MoS2 circuits with E-mode FETs for large-area electronics. Nano Lett. 2016, 16, 6349–6356. [Google Scholar] [CrossRef]
  8. Guo, Y.; Wei, X.; Shu, J.; Liu, B.; Yin, J.; Guan, C.; Han, Y.; Gao, S.; Chen, Q. Charge trapping at the MoS2-SiO2 interface and its effects on the characteristics of MoS2 metal-oxide-semiconductor field effect transistors. Appl. Phys. Lett. 2015, 106, 103109. [Google Scholar] [CrossRef]
  9. Ionescu, A.M.; Riel, H. Tunnel field-effect transistors as energy-efficient electronic switches. Nature 2011, 479, 329–337. [Google Scholar] [CrossRef]
  10. Sicard, E. Introducing 7-nm FinFET Technology in Microwind; HAL: Lyon, France, 2017. [Google Scholar]
  11. Peng, S.; Jin, Z.; Yao, Y.; Li, L.; Zhang, D.; Shi, J.; Huang, X.; Niu, J.; Zhang, Y.; Yu, G. Metal-contact-induced transition of electrical transport in monolayer MoS2: From thermally activated to variable-range hopping. Adv. Electron. Mater. 2019, 5, 1900042. [Google Scholar] [CrossRef]
  12. Zou, X.; Xu, J.; Huang, H.; Zhu, Z.; Wang, H.; Li, B.; Liao, L.; Fang, G. A comparative study on top-gated and bottom-gated multilayer MoS2 transistors with gate stacked dielectric of Al2O3/HfO2. Nanotechnology 2018, 29, 245201. [Google Scholar] [CrossRef]
  13. Ayari, A.; Cobas, E.; Ogundadegbe, O.; Fuhrer, M.S. Realization and electrical characterization of ultrathin crystals of layered transition-metal dichalcogenides. J. Appl. Phys. 2007, 101, 014507. [Google Scholar] [CrossRef] [Green Version]
  14. 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] [Green Version]
  15. Gray, P.V.; Brown, D.M. Density of SiO2–Si interface states. Appl. Phys. Lett. 1966, 8, 31–33. [Google Scholar] [CrossRef]
  16. Lu, C.P.; Li, G.; Mao, J.; Wang, L.M.; Andrei, E.Y. Bandgap, mid-gap states, and gating effects in MoS2. Nano. Lett. 2014, 14, 4628–6433. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  17. Lee, J.-M.; Cho, I.-T.; Lee, J.-H.; Kwon, H.-I. Bias-stress-induced stretched-exponential time dependence of threshold voltage shift in InGaZnO thin film transistors. Appl. Phys. Lett. 2008, 93, 093504. [Google Scholar] [CrossRef]
  18. Illarionov, Y.Y.; Knobloch, T.; Waltl, M.; Rzepa, G.; Pospischil, A.; Polyushkin, D.K.; Furchi, M.M.; Mueller, T.; Grasser, T. Energetic mapping of oxide traps in MoS2 field-effect transistors. 2D Mater. 2017, 4, 025108. [Google Scholar] [CrossRef]
  19. Miczek, M.; Mizue, C.; Hashizume, T.; Adamowicz, B. Effects of interface states and temperature on the C-V behavior of metal/insulator/AlGaN/GaN heterostructure capacitors. J. Appl. Phys. 2008, 103, 104510. [Google Scholar] [CrossRef] [Green Version]
  20. Yang, Z.; Liu, X.; Zou, X.; Wang, J.; Ma, C.; Jiang, C.; Ho, J.C.; Pan, C.; Xiao, X.; Xiong, J. Performance limits of the self-aligned nanowire top-gated MoS2 transistors. Adv. Funct. Mater. 2017, 27, 1602250. [Google Scholar] [CrossRef]
  21. Deuling, H.; Klausmann, E.; Goetzberger, A. Interface states in Si-SiO2 interfaces. Solid State Electron. 1972, 15, 559–571. [Google Scholar] [CrossRef]
  22. Park, Y.; Baac, H.W.; Heo, J.; Yoo, G. Thermally activated trap charges responsible for hysteresis in multilayer MoS2 field-effect transistors. Appl. Phys. Lett. 2016, 108, 083102. [Google Scholar] [CrossRef]
  23. Yang, G.; Chuai, X.; Niu, J.; Wang, J.; Shi, X.; Wu, Q.; Su, Y.; Zhao, Y.; Liu, D.; Xu, G. Anomalous positive bias stress instability in MoS2 transistors with high-hydrogen-concentration SiO2 gate dielectrics. IEEE Electron. Device Lett. 2018, 40, 232–235. [Google Scholar] [CrossRef]
Figure 1. (a) Schematic of the fabricated MoS2 FET. (b) SEM image of the MoS2 FET. The drain and sources are indicated by dashed boxes. (c,d) AFM profile revealing the MoS2 thickness along the red line.
Figure 1. (a) Schematic of the fabricated MoS2 FET. (b) SEM image of the MoS2 FET. The drain and sources are indicated by dashed boxes. (c,d) AFM profile revealing the MoS2 thickness along the red line.
Materials 13 02896 g001
Figure 2. (a) The transfer curves for forward sweep (red line) and backward sweep (black line) measured at 310 K. The hysteresis is selected as the voltage shift between the transfer curves of the forward sweep and the backward sweep at IDS = 100 pA, which is indicated by a double-headed arrow. The inset shows the positions of the Fermi level corresponding to VBG, mins of −50 to −80 V, respectively. (b,c) The band structure of MoS2 channel considering the interface states at the MoS2–SiO2 interface.
Figure 2. (a) The transfer curves for forward sweep (red line) and backward sweep (black line) measured at 310 K. The hysteresis is selected as the voltage shift between the transfer curves of the forward sweep and the backward sweep at IDS = 100 pA, which is indicated by a double-headed arrow. The inset shows the positions of the Fermi level corresponding to VBG, mins of −50 to −80 V, respectively. (b,c) The band structure of MoS2 channel considering the interface states at the MoS2–SiO2 interface.
Materials 13 02896 g002
Figure 3. Transfer curves at a series of temperatures (10–310 K) for forward sweep (a) and backward sweep (b), respectively. Relationship between VTH, VBG, min, and temperature for forward sweep (c) and backward sweep (d), respectively.
Figure 3. Transfer curves at a series of temperatures (10–310 K) for forward sweep (a) and backward sweep (b), respectively. Relationship between VTH, VBG, min, and temperature for forward sweep (c) and backward sweep (d), respectively.
Materials 13 02896 g003
Figure 4. Relationship between SS and temperature for forward sweep (a) and backward sweep (b), respectively. (c) Schematic of Fermi level shifting progress of MoS2 channel relative to the interface states and conduction band caused by the increasing of temperature from 10 to 310 K for VBG, min of −50 V. EF, +50 V and EF, sub correspond to the Fermi level for gate voltage of +50 V and subthreshold voltage, respectively.
Figure 4. Relationship between SS and temperature for forward sweep (a) and backward sweep (b), respectively. (c) Schematic of Fermi level shifting progress of MoS2 channel relative to the interface states and conduction band caused by the increasing of temperature from 10 to 310 K for VBG, min of −50 V. EF, +50 V and EF, sub correspond to the Fermi level for gate voltage of +50 V and subthreshold voltage, respectively.
Materials 13 02896 g004

Share and Cite

MDPI and ACS Style

Huang, X.; Yao, Y.; Peng, S.; Zhang, D.; Shi, J.; Jin, Z. Effects of Charge Trapping at the MoS2–SiO2 Interface on the Stability of Subthreshold Swing of MoS2 Field Effect Transistors. Materials 2020, 13, 2896. https://doi.org/10.3390/ma13132896

AMA Style

Huang X, Yao Y, Peng S, Zhang D, Shi J, Jin Z. Effects of Charge Trapping at the MoS2–SiO2 Interface on the Stability of Subthreshold Swing of MoS2 Field Effect Transistors. Materials. 2020; 13(13):2896. https://doi.org/10.3390/ma13132896

Chicago/Turabian Style

Huang, Xinnan, Yao Yao, Songang Peng, Dayong Zhang, Jingyuan Shi, and Zhi Jin. 2020. "Effects of Charge Trapping at the MoS2–SiO2 Interface on the Stability of Subthreshold Swing of MoS2 Field Effect Transistors" Materials 13, no. 13: 2896. https://doi.org/10.3390/ma13132896

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