Introduction

Ever since the inception of metal oxide semiconductor field effect transistor (MOSFET), Scaling has been the primary driving force behind its unprecedented success. The early era of scaling (1975–2005: Dennard Scaling) had two characteristic features: dimension scaling which allowed the number of transistor per chip to increase by 1000000x and consequently their speed to increase by 1000x, and voltage scaling which kept the power density practically constant throughout this scaling regime1,2. However, around 2005, the voltage scaling almost stopped as further reduction in the supply voltage (VDD) and hence the threshold voltage (VTH) was leading to exponential increase in the OFF state current (IOFF)3. This is a direct consequence of non-scalability of the subthreshold swing (SS) to below 60 mV/decade arising out of Boltzmann statistics that governs the operation of conventional MOSFETs. Dimension scaling, however, continued beyond 2005, but, under the new generalized scaling rules. This inevitably led to increase in the power density at the same rate as the integration density. The actual scenario is made worse by non-scaling factors which escalated static and leakage power densities at a much faster rate3. Power/heat dissipation, henceforth, became the main problem for high performance microprocessors. Today, in 2016, even dimension scaling seems extremely challenging beyond 10 nm gate length (LG) owing to fundamental material limitations. So it is not too far when all aspects of MOSFET scaling will completely stop, marking the end of the silicon complementary metal oxide semiconductor (CMOS) era. Therefore, in order to restore the golden era of transistor scaling, energy efficient and high performance innovative device ideas based on aggressively scalable novel materials need to be conceived on an urgent and immediate basis.

From the above discussion it is obvious that post-Si-CMOS devices have to resolve two key challenges: length scaling and voltage scaling. For length scaling, low dimensional systems like nanotubes4,5, nanowires6,7 and very recently nanosheets8,9,10,11,12,13,14 are being considered as alternative materials to Si due to their inherent electrostatic integrity that allows fundamentally superior scaling properties. Voltage scaling, however, necessitates steep slope devices which in turn require operation beyond Boltzmann statistics. Several steep switching device concepts like tunneling FETs15,16, piezoelectric strain modulated Si FinFETs17, negative capacitance ferroelectric FETs18 excitonic FETs19 and spin-based FETs20,21 have been proposed. Among these, tunneling FETs are the most matured candidates which have experimentally demonstrated SS less than 60 mV/decade15. However, the greatest challenge for tunneling FETs are their low ON state current densities limited by large tunneling barriers. Piezoelectric strain modulated Si FinFETs are also promising but suffers from the limitation of bulk nature of Si at the scaling limits. The readers should note that the objective for post-CMOS device design is mainly two-fold: SS slope should be as abrupt as possible (ideally zero) to meet the low power requirement whereas the ON current should be as high as possible to increase device speed.

In this context, 2D layered semiconductors are receiving significant attention as possible candidates for post-Si electronics owing to their ultra-thin body nature that allows aggressive channel length scaling and hence high performance. Moreover, recent experimental and theoretical studies show that the bandgap of multilayer transition metal dichalcogenides (TMDs: an extensively studied class of 2D materials) like MoS2, WSe2 etc. can be dynamically reduced to zero by applying out-of-plane stress22,23. In this article we, therefore, combine the superior scalability of 2D materials with stress induced dynamic bandgap engineering into a breakthrough device concept called Two Dimensional Electrostrictive Field Effect Transistor or 2D-EFET. As shown in Fig. 1, the device structure resembles a conventional 2D-FET with the exception that the gate insulator is substituted with an electrostrictive material. The operating principal, however, differs significantly since an electrostrictive material not only behaves like a high-k insulating gate oxide but also undergoes longitudinal expansion when an electric field is applied across it. This expansion transduces an out-of-plane stress on the 2D channel material and monotonically reduces its bandgap to zero. In the OFF state, the 2D-EFET operates with a large bandgap and prevents current conduction whereas in the ON state, it operates with a smaller or zero bandgap and allows current conduction. The fascinating aspect of 2D-EFET is that it offers steep SS below 60 mV/decade owing to an internal feedback mechanism giving rise to voltage amplification and at the same time provides significantly higher ON state current density compared to any existing charge based device concepts.

Figure 1
figure 1

Proposed 2D-EFET.

In the OFF state the 2D channel material offers a large bandgap and restricts current flow whereas in the ON state the 2D channel material offers a small or zero bandgap and allows current flow. The expansion in the electrostrictive material in response to the applied gate bias is compensated by reduction in inter-layer distance between two consecutive layers of the 2D material. This transduction results in an internal voltage amplification which allows the SS to become less than 60 mV/decade.

The operation of 2D-EFET is conceptually explained in Fig. 2 and through equation 1 through 8. Figure 2a shows the strain (S) versus electric field (ξ) characteristics (equation 1) of an electrostrictive material and Fig. 2b shows the bandgap (EG) versus out-of-plane stress (P) characteristics (equation 2) of a 2D material based on phenomenological models. The parameter sets (α, β, δ) and (EG0, κ, χ) are specific to the choice of electrostrictive and 2D material and their values used for our simulations were extracted from experimental literature (the linear approximation of equation 1 and 2 which are also shown in Fig. 2a,b will be used primarily for easier qualitative discussion). In 2D-EFET, equations 1 and 2 are coupled through equation 3, where, tE is the thickness of the electrostrictive material, ϒ2D and t2D are, respectively, the Young’s modulus and thickness of the 2D material, and finally, η quantifies the ratio of strain transfer from the electrostrictive material to the channel material such that Δt2D = η.ΔtE (0 < η < 1). Note that η is a very important parameter for the 2D-EFET since it determines the efficiency of strain transduction and hence the improvement in SS. It remains to be experimentally demonstrated how electrostrictive material can be integrated with 2D materials for maximum strain transfer. One possible strategy would be capping/encapsulating the entire device structure in a rigid fixture to minimize the expansion of the electrostrictive material away from the 2D channel. Other possible approach will be engineering better electrostrictive material which can provide higher strain than PMN-PT and thereby compensate for poor coupling efficiency. Recently, there has been significant progress towards the development of lead-free ceramics by various doping and alloying routes which includes titanates, alkaline niobates and bismuth perovskites and their solid solutions24. Specifically, giant electric field-induced strain (EFIS) was observed in Nb modified lead-free Bi0.5Na0.5TiO3 - Bi0.5K0.5TiO3 - LiTaO3 (BNT-BKT-LT) ternary system. With 3 mol% Nb substitution, the EFIS could be enhanced up to 641 pm/V25. Lead-free Zr-modified Bi0.5(Na0.78K0.22)0.5TiO3 ceramics (BNKTZ-100x, with x = 0–0.05) also demonstrated enhanced unipolar field-induced strain of 0.43%26.

Figure 2
figure 2

Operating principle for 2D-EFET.

(a) Strain (S) versus electric field (ξe) characteristics of an electrostrictive material (PMN-PT). (b) Bandgap (EG) versus out-of-plane stress (P) characteristics of a 2D material (MoS2). (c) Equivalent capacitive network for a 2D-EFET. (d) Schematic showing the band movement in 2D-EFET in response to applied gate bias (VGS). ΨS is the usual electrostatic component, whereas ΨE is the electrostrictive component. ΨE arises due to the reduction in the bandgap of the 2D material in response to the out-of-plane stress transduced by the electrostrictive material. This gives rise to internal voltage amplification which is ultimately responsible for steep subthreshold swing (SS) less than 60 mV/decade in 2D-EFETs.

Figure 2c shows an equivalent capacitive network model for our 2D-EFET where, VGS is the applied external gate bias, ΨS is the electrostatic surface potential and CE, CCH and CIT are respectively the capacitances associated with the electrostrictive material, the 2D channel material and the interface traps. Figure 2d shows position of the energy bands inside the 2D channel material corresponding to the OFF and ON state of device operation. Note that the additional band movement ΨE appears due to the decrease in the bandgap of the channel material through electrostrictive transduction. Theoretical simulations performed on bilayer TMDs by Kumar et al.23 suggests that the conduction band (CB) minima between the K-Γ high symmetry point and the valence band (VB) maxima at the Γ point in the Brillouin zone of various TMDs move towards their corresponding Fermi levels when the inter-layer spacing between the successive van der Waals layers are monotonically reduced. It is also apparent from their calculations that the rate of movements of CB minima and VB maxima as a function of the inter-layer spacing are almost similar which justifies the assumption ΨE = 1/2   ΔEG. Note that the total channel potential (ΨT = ΨS + ΨE) is always greater than ΨS and results in an internal voltage amplification. Figure 3 shows electrostatic (ΨS), electrostrictive (ΨE) and total channel potential (ΨT) as a function of external gate bias (VGS) obtained by solving equation 4 through 6 self-consistently with equation 7. Note that CCH enters into the expression for ΨS and ΨT through the quantity r. In equation 7, D(E) denotes the 2D density of states derived from parabolic energy dispersion relationship, m* is the carrier effective mass, h is the Planck’s constant and fS(E) and fD(E) are the Fermi function for the source and drain contact electrodes, respectively. Finally equation 8 (kB is the Boltzmann constant, q is the electronic charge and T is the temperature) represents the subthreshold swing (SS).

Figure 3
figure 3

Channel Potential Map.

(a) Electrostatic (ΨS), (b) Electrostrictive (ΨE) and (c) Total channel potential (ΨT) as a function of applied gate bias (VGS) for γ = 0.0 (black), γ = 0.7 (red), γ = 1.5 (blue) and γ = 2.0 (green) with CIT = 0. Clearly, the electrostatic band movement is one to one (r = 1) with VGS in the subthreshold regime (OFF state) of the device operation and as such no electrostrictive transduction occurs. However, as the device enters near threshold or ON state operation feedback mechanism begins between electrostatic and electrostrictive potentials since r < 1 owing to finite channel capacitance (CCH). This leads to internal voltage amplification (ΨT > VGS) in the ON state as shown in Fig. 3c. The dotted black line represents ΨT = VGS. However, this amplification is not conducive for sub-60 mV/decade SS. Figure (c–e) shows respectively ΨS, ΨE and ΨT as a function of VGS for same values of γ but with finite CIT = 3CE. In these instances, electrostrictive transduction take place even in the OFF state. However, internal voltage amplification only occurs with γ = 1.5 (blue) and γ = 2.0 (green). The material parameters used for our self-consistent simulations are as follows: α = 8e-5 (cm/kV)2, β = 20, δ = 3.5(cm/kV)1.5, EG0 = 1.6 eV, κ = 70 meV/GPa, χ = 1.2 meV/(GPa)2, ϒ2D = 300GPa, t2D = 1 nm, tE = 100 nm, m* = 0.45m0. The values of η (0 < η < 1) were adjusted to obtain the different γ values.

As evident from equation 8, in order to obtain sub-60 mV/decade SS, it is necessary to make ∂ΨT/∂VGS > 1 in the OFF state of the device operation. This is possible if and only if the quantity r < 1 which could be achieved through a finite value of CCH or CIT or both. Now, in a conventional 2D-FET (where γ = 0 and, therefore, ΨE = 0 and ΨT = ΨS), CCH is negligible in the OFF state since D(E) ≈ 0 and CIT should ideally be negligible so that r = 1 (i.e. ΨS = VGS, in equation 4) and hence SS = 60 mV/decade. This is the best possible SS achievable within Boltzmann limit. In 2D-FETs r < 1 leads to SS > 60 mV/decade which is undesirable. But, this scenario is completely different for the proposed 2D-EFET, since r = 1 results in ξ = 0 and hence ΨE = 0. So no transduction leading to internal voltage amplification can occur. As shown in Fig. 3a, if CIT = 0, the electrostatic band movement is almost one to one with the applied gate bias in the OFF state of the device (VGS < VFB − vTB, where, VFB is the flat band voltage and vTB denotes the thermal broadening; vTB = 6kBT. We have used VFB = 0.4 V for our simulations). This is true irrespective of the value of γ. However, as the device enters near threshold or ON state operation, CCH becomes finite leading to r < 1 and ΨE > 0 as shown in Fig. 3a,b respectively. This initiates an internal feedback mechanism which requires a readjustment of ΨS since CCH depends on both electrostatic and electrostrictive potential (equation 7). Hence, ΨS becomes non-monotonic in the ON state of device operation as shown in Fig. 3a. Note that higher values of γ are associated with more efficient strain transduction and hence stronger and more abrupt feedback mechanism. The total channel potential ΨT, however, increases monotonically according to Fig. 3c. But, none of these scenarios are conducive for achieving sub-60 mV/decade SS since the internal voltage amplification (ΨT > ΨS) takes place only in the ON state of the device operation. As shown in Fig. 3d–f, more favorable conditions can be achieved when CIT is finite. This allows r < 1 and hence ΨE > 0 in the OFF state of the device operation which ultimately leads to internal voltage amplification (ΨT > ΨS). However, the criterion for sub-60 mV/decade SS i.e. ∂ΨT/∂VGS > 1 is achieved only when γ > 1 and r < 1 as obvious from equation 9. Note that internal voltage amplification leading to sub-60 mV/decade SS can also be achieved in negative capacitance ferroelectric FETs (NCFETs). However, the working principle of NCFET is distinctly different compared to 2D-EFET18.

Figure 4a,b shows the transfer characteristics of 2D-EFET obtained by solving equations 4 through 7 self consistently with ballistic Landauer formalism27 (equation 9).

Figure 4
figure 4

2D-EFET Device Characteristics.

Room temperature current (IDS) versus gate voltage (VGS) characteristics of 2D-EFET for γ = 0.0 (black), γ = 0.7 (red), γ = 1.5 (blue) and γ = 2.0 (green) with (a) CIT = 0 and (b) CIT = 3CE. The η values used to obtain the corresponding γ values are 0.0, 0.2, 0.4, and 0.6, respectively. Source to drain bias of VDS = 0.1 V and flat band voltage of VFB = 0.4 V was used. (c) Subthreshold Swing (SS) as a function of γ for different values of CIT. Clearly sub-60 mV/decade SS is obtained if and only if CIT > 0 and γ > 1. (d) Dynamic bandgap change in the channel of a 2D-EFET due to electrostrictive transduction for γ = 2, CIT = 3CE and VFB = 0.4 V. (e) Room temperature output characteristics (IDS versus VDS) for the same 2D-EFET for different values of VGS. The rectangular boxes in figure (d,e) correlate the length of the saturation region with the dynamic bandgap of the 2D-EFET. (f) ON current (ION) and ON to OFF current ratio (ION/IOFF) as a function of VFB for different supply voltages (VDS = VGS = VDD). Ballistic Landauer formalism was used to compute the current versus voltage characteristics.

In equation 9, I1 and I2 are the current due to electron injection from the drain and the source contacts, respectively and v(E) is the carrier velocity. As expected sub-60 mV/decade SS is achieved for finite value of CIT and γ > 1. Figure 4c shows the average SS (over 4 decades) as a function of γ for different values of CIT. It is counter intuitive from a conventional 2D-FET stand point to note that higher values of CIT allow better subthreshold slopes in 2D-EFET for a given value of γ. But the reader should realize that although finite CIT is detrimental for electrostatic band movement, it is beneficial for electrostrictive band movement. This is because larger CIT ensures larger potential drop across the electrostrictive gate material which culminates into higher stress and greater bandgap reduction in the 2D channel material. In principle, if we make r = 0 corresponding to an infinitely large CIT, then ΨS = 0, which completely stops the electrostatic band movement in the channel. However, ΨE will still be finite due to the electrostrictive band movement. As long as ΔΨE > ΔVGS, internal voltage amplification will occur leading to SS < 60 mV/decade. Note that the improvement in the SS in 2D-EFET is essentially determined through the quantity γ which includes several material parameters corresponding to both the 2D material and the electrostrictive material. A larger overall value for γ will ensure a better SS as evident from equation 8. A thicker channel material will certainly make γ small in accordance with equation 6, whereas monolayer materials obviously lack van der Waals spacing. Bilayer materials are optimum for 2D-EFET since these are the thinnest with a finite van der Waals spacing. In the manuscript we have used the material parameters corresponding to MoS2 since it has been most extensively studied in experiments. Other TMDs also demonstrate similar effects, however there is lack of experimental literature. In fact as shown by Kumar et al.23, similar bandgap changes can be obtained at smaller strain values for MoSe2 and MoTe2 i.e. these materials offer larger κ values and hence larger γ.

Figure 4d shows the dynamic bandgap change as a function of the applied gate bias and Fig. 4e shows the output characteristics for the 2D-EFET with γ = 2.0 and CIT = 3CE. The reader should remember that the length of the plateau in the output characteristics which is referred to as the saturation regime is equal to the bandgap of the channel material (at T = 0 K and shortened by few kBT due to thermal broadening at finite T) for a ballistic transistor. Since the bandgap changes in our 2D-EFET as a function of the applied gate bias, the length of the saturation region also changes accordingly in Fig. 4e. Figure 4f shows the ON current (ION) and ON to OFF current ratio (ION/IOFF) for the 2D-EFET as a function of the flat band voltage (VFB) for different supply voltages (VDD). As expected ION increases with increase in VDD but decreases with increase in VFB, whereas, IOFF remains practically constant with VDD and decreases exponentially with increase in VFB resulting in the trends observed in Fig. 4f. Clearly, the standard requirement for high performance FETs (i.e. ION = 1 mA/μm and ION/IOFF = 104) can be achieved by 2D-EFET for VDD = 0.2 V which is 3 times smaller than the predicted VDD = 0.6 V for 2020 by ITRS. Moreover, ON current as high as 10 mA/μm can be delivered by 2D-EFET for a supply voltage of VDD = 0.4 V. One criticism could arise from achievable operating speed in 2D-EFET since mechanical motions are slow and limited by the acoustic velocities. However, a simple back of the envelop calculation (conservative) with tE = 100 nm and ve = speed of sound in solid = 1000 m/s results in ~100 GHz operating speed. One obvious way to increase the speed would be the scaling of the thickness of the electrostrictive material (tE), however, this may lead to reduction in strain transfer and hence increase in SS. Optimization will be investigated in future studies. Finally, channel length (LG) scalability of 2D-EFET is determined through the band bending length λ (equation 10), which is derived by solving 2D Poisson’s equation similar to conventional planar FETs6. Therefore, from a pure electrostatic point of view, the use of ultra-thin 2D channels (t2D = 1 nm, ε2D = 8), with ultra-high-k dielectric (tE = 100 nm, εE = 2000) allows scalability (LG > 3λ = 1.8 nm) well beyond 10 nm technology node6.

In the conclusion, a novel and disruptive device concept called 2D-EFET based on strain transduction and dynamic bandgap engineering in 2D material has been proposed and numerically evaluated. The 2D-EFET provides a breakthrough solution for post silicon, ultra-low power, high performance and aggressively scalable device technology.

Additional Information

How to cite this article: Das, S. Two Dimensional Electrostrictive Field Effect Transistor (2D-EFET): A sub-60mV/decade Steep Slope Device with High ON current. Sci. Rep. 6, 34811; doi: 10.1038/srep34811 (2016).