The present description relates generally to particle tunneling and field effect transistors and, more particularly, to two-dimensional heterojunction interlayer tunneling field effect transistors.
Electronic integrated circuits may be considered the hardware backbone of today's information society. However, power dissipation of such circuits has recently become a considerable challenge. Rates of power consumption in these integrated circuits can affect, for example, the useful lifespan of portable equipment, the sustainability of the ever-increasing number of large data centers, the feasibility of energy-autonomous systems in terms of ambience intelligence, and the feasibility of sensor networks associated with implants and other medical devices, among others. While the scaling of a supply voltage (VDD) is recognized as one of the most effective measures for reducing switching power in digital circuits, the performance loss and increased device-to-device variability are typically seen as serious hindrances to scaling VDD down to 0.5 volts (V) or less.
As the physical limitations of miniaturization appear to approach for complementary metal-oxide-semiconductor (CMOS) technology, the search for alternative devices to extend computer performance has accelerated. In general, any new technology should be energy efficient, dense, and enable more device function per unit space and time. There have been many device proposals, often involving new state variables and communication frameworks. Moreover, it is known in the art that the voltage scalability of very-large-scale integration (VLSI) systems may be significantly improved by resorting to innovations in transistor technology and, in this regard, the International Technology Roadmap for Semiconductors (ITRS) has singled out tunnel field effect transistors (“TFETs” or “tunnel FETs”) as the most promising transistors to reduce sub-threshold swing (SS) below the 60 mV/dec limit of metal-oxide-semiconductor field-effect transistors (MOSFETs) at room temperature and, thus, to enable further VDD scaling. Several device architectures and materials are being investigated to develop tunnel FETs offering both an attractive on-current and a small SS, including group III-group V based transistors, possibly employing staggered or broken bandgap heterojunctions, or strain engineering. Even if encouraging experimental results have been reported for the on-current in group III-V tunnel FETs, achieving a sub-60 mV/dec SS remains a major challenge in these devices, likely due to the detrimental effects of interface states. Therefore, as of now, the investigation of new material systems and innovative device architectures for high performance tunnel FETs is as timely as ever in both the applied physics and the electron device community.
The following description of example methods and apparatus is not intended to limit the scope of the description to the precise form or forms detailed herein. Instead the following description is intended to be illustrative so that others may follow its teachings.
Monolayers of group-VIB transition metal dichalcogenides (TMDs) according to the formula MX2—where M=Mo or W, and where X=S, Se, or Te—have recently attracted attention for their electronic and optical properties. As explained below, these materials may be utilized by the 2D crystal layers in the example 2D heterojunction interlayer tunneling field effect transistors (Thin-TFETs) disclosed herein. Monolayers of TMDs have a bandgap that varies from almost zero to 2 eV with a sub-nanometer thickness. As a result, these materials are considered to be approximately two-dimensional (2D) crystals. 2D crystals, in turn, have recently attracted attention primarily due to their scalability, step-like density of states, and absence of broken bonds at interface. 2D crystals can be stacked to form a new class of tunneling transistors based on an interlayer tunneling occurring in the direction normal to the plane of the 2D materials. In fact, tunneling and resonant tunneling devices have recently been proposed, as well as experimentally demonstrated for graphene-based transistors.
Further, the sub-nanometer thickness of TMDs provides excellent electrostatic control in a vertically stacked heterojunction. What's more, the 2D nature of such materials makes them virtually immune to the energy bandgap increase produced by the vertical quantization when conventional 3D semiconductors are thinned to a nanoscale thickness and, thus, immune to the corresponding degradation of the tunneling current density. Still further, the lack of dangling bonds at the surface of TMDs may allow for the fabrication of material stacks with low densities of interface defects, which is another potential advantage of TMD materials for tunnel FET applications.
With reference now to the figures,
Further, in some examples the Thin-TFET 100 includes the interlayer 118, which separates the top and bottom 2D layers 106, 112. The interlayer 118 may, in some cases, take the form of a van der Waals gap that is formed by the lack of chemical bonds between the top and bottom 2D layers 106, 112. Of course, the Thin-TFET 100 is not in any way limited to only those examples in which not a single chemical bond is present between the top and bottom 2D layers 106, 112. As those having ordinary skill in the art will understand, in some examples at least some chemical bonds may be present between the top and bottom 2D layers 106, 112 of the Thin-TFET 100. However, in some instances, material selection of the top and bottom 2D layers 106, 112 is important so as to prevent, or at least minimize, such chemical bonds at the interlayer 118. The example top and bottom 2D layers 106, 112 may be atomically-thick monolayer 2D crystals whose surfaces are free, or at least substantially free, from dangling bonds. Hence, even though
Furthermore, it should also be understood that references to “top” and “back”/“bottom” herein may in some examples be interchangeable with references to “first” and “second,” respectively and do not necessarily indicate a required orientation of the Thin-TFET 100, but rather are used merely to assist in understanding the structure of the device. Still further, while the top gate 102, the top oxide layer 104, the top 2D layer 106, the interlayer 118, the bottom 2D layer 112, the back oxide layer 110, and the back gate 108 are aligned in a vertically stack (or “configuration”) in
With reference now to
To determine the band alignment in a vertical direction of the example Thin-TFET 100 in
C
TOX
V
TOX
−C
IOX
V
IOX
=e(pT−nT+ND),
C
BOX
V
BOX
−C
IOX
V
IOX
=e(pB−nB+NA), (1)
where CTOX, CIOX, and CBOX are the capacitances per unit area of, respectively, the top oxide layer 104, the interlayer 118, and the back oxide layer 110 and where VTOX, VIOX, and VBOX are the potential drops across, respectively, the top oxide layer 104, the interlayer 118, and the back oxide layer 110. In one example, the potential drop across the top and back oxide layers 104, 110 can be written in terms of the external voltages VTG, VBG, VDS, and in terms of the energy eΦn.T=ECT−EFT and eΦp.T=EFB−EVB defined in
eV
TOX
=eV
TG
+eφ
n.T
−eV
DS+χ2D.T−ΦM.T.,
eV
BOX
=eV
BG
−eφ
p.B
+E
GB+χ2D.B+ΦM.B.,
eV
IOX
=eV
DS
+eφ
p.B
−eφ
n.T
+E
GB+χ2D.B−χ2D.T (2)
where EFT and EFB are Fermi levels of majority carriers in the top and bottom 2D layers 106, 112. In some examples, nT, pT are the electron and hole concentrations in the top 2D layer 106; nB, pB are the concentrations in the bottom 2D layer 112; x2D,T, x2D,B are the electron affinities of the top and bottom 2D layers 106, 112; ΦT and ΦB are the work functions of the top and back gates 102, 108; and EGB is the energy gap in the bottom 2D layer 112. Equation (2) is based on an assumption that majority carriers of the top and bottom 2D layers 106, 112 are at thermodynamic equilibrium with their Fermi levels, with the split of the Fermi levels set by the external voltages (i.e., EFB−EFT=eVDS), and the electrostatic potential essentially constant in the top and bottom 2D layers 106, 112.
Because a parabolic effective mass approximation for the energy dispersion of the 2D materials is employed herein, the carrier densities can be expressed as an analytic function of eΦn.T and eΦp.B
where gv is the valley degeneracy.
In some examples it is possible to determine the tunneling current of the example Thin-TFET 100 based on the transfer-Hamiltonian method used in the context of resonant tunneling in graphene transistors. The single particle elastic tunneling current may be represented as
where e is the elementary charge; kB and kT are wave-vectors, respectively, in the bottom and top 2D layers 112, 106; where EB(kB) and ET(kT) denote corresponding energies of the bottom and top 2D layers 112, 106; where fB and fT are Fermi occupation functions in the bottom and top 2D layers 112, 106 (i.e., depending respectively on EFB and EFT with respect to
M(kT,kB,)=∫Adr∫dzT,k
where ΨB,k
In some examples, to determine M (kT, kB), the electron wave-function may be written in Bloch function form as
where uk (r, z) is a periodic function of r and where Nc is the number of unit cells in an overlapping area A of the top and bottom 2D layers 106, 112. Equation (6) assumes the following normalization condition:
∫Ω
where ρ is the in-plane abscissa in the unit cell area ΩC and the overlapping area A=NCΩC.
The wave-function Ψk (r, z) presumably decays exponentially in the interlayer 118 with a decay constant κ. Such a z dependence can be absorbed in uk(r, z) based on various derivations as will be understood by those having ordinary skill in the art. Moreover, it should be understood that absorbing the exponential decay in uk (r, z) accounts for the fact that in the interlayer 118 the r dependence of the wave-function changes with z in some instances. In fact, as disclosed above, while uk (r, z) is localized around basis atoms in the top and bottom 2D layers 106, 112, these functions spread out while they decay in the interlayer 118 so that the r dependence becomes weaker when moving farther from the 2D layers.
To determine M (kT, kB), a scattering potential in the interlayer 118 may be separable in the form
U
sc(r,z)=VB(z)FL(r), (8)
where FL(r) is the in-plane fluctuation of the scattering potential, which is essentially responsible for the relaxation of momentum conservation in the tunneling process. By substituting Equations (6) and (8) into Equation (5) and writing r=rj+ρ, where rj is a direct lattice vector and ρ is the in-plane position inside each unit cell, the following is obtained:
In some cases, FL(r) corresponds to relatively long range fluctuations, so that FL(r) is relatively constant inside a unit cell and that, furthermore, the top and bottom 2D layers 106, 112 have the same lattice constant. Hence the Bloch functions uT.k
where the integral in the unit cell has been written for rj=0 because it is independent of the unit cell.
In keeping with kB and kT being small compared to the size of the Brillouin zone, in Equation 10 the kB (kT) dependence of uB,k
∫Ω
where TIL represents a thickness of the interlayer 118 and MB0 is a k independent matrix element that remains a prefactor in the final expression for the tunneling current. Because FL(r) is a slowly varying function over a unit cell, the sum over the unit cells in Equation (10) can be rewritten as a normalized integral over the tunneling area A
Still further, by introducing Equations (11) and (12) into Equation (10), the squared matrix element can be represented as
where q=kB−kT and where SF(q) is a power spectrum of the random fluctuation described by FL(r), which is defined as
Yet further, by substituting Equation (13) into Equation (4) and then converting the sums over kB and kT to integrals, the following is obtained:
According to Equation (15), current is proportional to the squared matrix element |MBO|2 defined in Equation (11) and decreases exponentially with the thickness TIL of the interlayer 118 according to the decay constant κ of the wave-functions. The equations thus far resort to a semi-empirical formulation of the matrix element given by Equation (11), where MBO is left as a parameter to be determined and discussed by comparing to experiments. A multitude of challenges are avoided by doing so. However, those having ordinary skill in the art would recognize how to modify the equations identified above if, for example, one were to derive a quantitative expression for MBO, if one were to specify how the periodic functions u0T(ρ, z) and u0B (ρ, z) spread out when they decay in the barrier region, and if one were to identify what potential energy and/or which Hamiltonian should be used to describe the barrier region itself (e.g., an effective barrier height of the van der Waals gap between two 2D crystals of 1.0 eV). Likewise, it should be understood that even though giant spin-orbit couplings have been reported in 2D TMDs, the effects of spin-orbit interaction in the bandstructure of 2D materials have been omitted from the equations above. Also, if energy separations between spin-up and spin-down bands are large, then the spin degeneracy in current calculations should be one instead of two, which would affect the magnitude, but not dependence on gate bias. Further, the equations above could also be modified to account for different band structures in TMD materials produced by a vertical electrical field. However, such effects are negligible due to the magnitude of the electrical field employed in the top and bottom 2D layers 106, 112 of the example Thin-TFET 100.
Nonetheless, in some examples the decay constant κ in the interlayer 118 may be approximated from the electron affinity difference between the top and bottom 2D layers 106, 112 and the interlayer 118 material. Moreover, according to Equation (15) the constant κ determines the dependence of the current on TIL, and κ in many cases is known according to prior studies (e.g., values of κ reported for an interlayer tunneling current in a graphene-hBN system).
As for the power spectrum SF(q) of the scattering potential, which is represented as
where q=|q| and where LC is the correlation length, which is assumed to be large compared to the size of a unit cell. In some instances, Equation (16) is consistent with an exponential form of an autocorrelation function of FL(r), and a similar q dependence is employed to reproduce the experimentally observed line-width of the resonance region in graphene interlayer tunneling transistors. Such a functional form is representative, at least in some examples, of phonon assisted tunneling, short-range disorder, charged impurities, or Moiré patterns (e.g., at a graphene-hBN interface). As explained below, the correlation length LC influences the gate voltage dependent current.
According to Equations (4) and (15), the tunneling current through the example Thin-TFET 100 is zero when there is no energy overlap between the conduction band ECT in the top 2D layer 106 and the valence band EVB in the bottom 2D layer 112 (i.e., ECT>EVB). It should be understood that the 2D materials of the top and bottom 2D layers 106, 112 inevitably have phonons, disorder, and host impurities and are affected by the background impurities in the surrounding materials. Hence a finite broadening of energy levels occurs because of the statistical potential fluctuations superimposed to the ideal crystal structure. The energy broadening in 3D semiconductors is known to lead to a tail of the density of states (DoS) in a gap region, which is also observed in optical absorption measurements and denoted the “Urbach tail.” It follows that in some examples the finite energy broadening is a fundamental limit to the abruptness of the turn on characteristic attainable with the example Thin-TFETs.
In some cases, energy broadening in 2D systems stems from interactions with randomly distributed impurities and disorder in the top and bottom 2D layers 106, 112 or in the surrounding materials, by scattering events induced by the interfaces, as well as by other scattering sources. For purposes of simplicity, a detailed description of energy broadening is omitted. Notwithstanding, the density of states ρ0(E) for a 2D layer with no energy broadening is
where E(k) denotes the energy relation with no broadening and where gs represents spin and where gv represents valley degeneracy. Put another way, in the presence of a randomly fluctuating potential V(r), the DoS can be written as
where Pv (v) is the distribution function for V(r), as explained below, and where the p0(E) definition in Equation (17) is used to go from the first equality to the second equality.
By way of comparison of Equation (18) to Equation (17), it can be seen that the ρ(E) of the example Thin-TFET 100 in the presence of energy broadening is calculated by substituting the Dirac function in Equation (17) with a finite width function Pv(v), which is the distribution function of V(r), and it is thus normalized to one.
To include the effects of energy broadening in the calculations, the tunneling rate is rewritten in Equation (4) as
It will be appreciated that, consistent with Equation (18), energy broadening can be included in the current calculation by substituting δ[E−E(k)] with Pv[E−E(k)]. In turn, the tunneling rate becomes
where an energy broadening spectrum SE is defined as
S
E(ET(kT)−EB(kB))=∫−∞∞dEPvT[E−ET(kT)]×PvB[E−EB(kB)] (21)
where PvT and PvB are potential distribution functions due to the presence of randomly fluctuating potential V(r) in, respectively, the top and the bottom 2D layers 106, 112.
In view of Equation (20), in terms of the tunneling current, energy broadening was accounted for by using in all calculations the broadening spectrum SE(ET(kT)−EB(kB)) defined in Equation (21) in place of δ[ET(kT)−EB(kB)]. More specifically, a Gaussian potential distribution was used for both the top and the bottom 2D layers 106, 112:
which has been derived for energy broadening induced by randomly distributed impurities, in which case σ is expressed in terms of the average impurity concentration.
Further, for the Gaussian spectrum in Equation (22), the overall broadening spectrum SE defined in Equation (21) is calculated analytically and reads
Hence SE also has a Gaussian spectrum, where σT and σB are, respectively, broadening energies for the top and bottom 2D layers 106, 112.
Many of the derivations above assumed a perfect rotational alignment between the lattice structures of the top and bottom 2D layers 106, 112 and that tunneling occurs between equivalent extrema in the Brillouin zone, that is, tunneling from a K to a K extremum (or from K′ to K′ extremum). As shown in
in terms of the rotational misalignment angle θ.
To be consistent, uT
By expressing r′ and k′ in the principal coordinate system, the matrix element can be written as
where q=(kB−kT) and the vector
Q
D
=K
0B
−{circumflex over (R)}
T→B
K
0T (26)
is introduced.
Equation (25) is an extension of Equation (10) and accounts for a possible rotational misalignment between the top and bottom 2D layers 106, 112 and also describes tunneling between nonequivalent extrema. The vector QD is zero only for tunneling between equivalent extrema (i.e., K0B=K0T) and for a perfect rotational alignment (i.e., θ=0). In a case where all extrema are at the K point and |K0B|=|K0T|=4π/3a0, then for K0B=KOT the magnitude of QD is given by QD=(8π/3a0)sin(θ/2).
One difference in Equation (25) compared to Equation (10) is that, in the presence of rotational misalignment, the top layer Bloch function u0T ({circumflex over (R)}B→Tr,z) has a different periodicity in the principal coordinate system from the bottom layer u0B (r, z). As a result, the integral over the unit cells of the bottom 2D layer 112 is not the same in all unit cells, so that the derivations going from Equation (10) to Equation (15) should be rewritten accounting for a matrix element MB0j depending on the unit cell j. Such an MB0j could be included in the calculations by defining a new scattering spectrum that includes not only the inherently random fluctuations of the potential FL(r), but also the cell to cell variations of the matrix element MB0j. A second difference of Equation (25) compared to Equation (10) lies in the presence of QD in the exponential term multiplying FL(rj).
In the case of tunneling between nonequivalent extrema and with a negligible rotational misalignment (i.e., θ≅0), Equation (26) gives QD=K0B K0T, and the current can be expressed as in Equation (15), but with the scattering spectrum evaluated at |q+QD|. Because in this case the magnitude of QD is comparable to the size of the Brillouin zone, the tunneling between nonequivalent extrema is substantially suppressed if the correlation length Lc of the scattering spectrum SF(q) is much larger than the lattice constant, as has been assumed in all derivations. Further, the derivations suggest that rotational misalignment affects the absolute value of the tunneling current, but not to change significantly its dependence on the terminal voltages.
Furthermore, if the vertical stack of the 2D materials is obtained using a dry transfer method, rotational misalignment is nearly inevitable. Tests have shown that, when the stack of 2D materials is obtained by growing the one material on top of the other, the top 2D layer 106 and the bottom 2D layer 112 have a fairly good angular alignment.
An analytical, approximated expression for the tunneling current is useful for a number of reasons, including to gain insight about the main physical and material parameters affecting the current versus voltage characteristic of the example Thin-TFET 100. To derive an analytical current expression, a parabolic energy relation is assumed, which allows for the following expression:
where EVB (kB), ECT(kT) are the energy relation, respectively, in the bottom 2D layer valence band and the top 2D layer conduction band and mv and mc are the corresponding effective masses.
It should be understood that energy broadening is neglected here, and Equation (15) is used as a starting point. Consequently, these equations are valid for the ON state of the example Thin-TFET 100 (i.e., ECT<EVB).
Turning to the integral over kB and kT in Equation (15) and introducing polar coordinates kB=(kB, θR), kT=(kT, θT) allows for the use of Equation (27) to convert the integrals over kB, kT to integrals over respectively EB, ET, which leads to
where the spectrum SF(q) is given by Equation (16) and thus depends only on the magnitude q of q=kB−kT. Assuming that ECT<EVB, the Dirac function reduces one of the integrals over the energies and sets E=EB=ET. Furthermore, the magnitude of q=kB−kT depends only on the angle θ=θB−θT, so that Equation (28) simplifies to
With respect to the ON state (i.e., ECT<EVB) for the example Thin-TFET 100, the zero Kelvin approximation for the Fermi-Dirac occupation functions fB, fT are introduced to further simplify Equation (29) to:
where Emin=max {ECT, EFT}, where Emax=min {EVB, EFB}, and where the tunneling window can be defined by [Emax−Emin].
The evaluation of Equation (30) requires expressing q as a function of the energy E inside the tunneling window and of the angle θ between kB and kT. Because q2=kB2+kT2−2kBkT cos(θ), Equation 27 can be written as follows:
where E=ER=ET. By substituting Equation (31) into the spectrum SF(q), the resulting integrals over E and θ in Equation (30) cannot be evaluated analytically. To proceed further, therefore, the maximum value taken by q2 is examined. The θ value leading to the largest q2 is θ=π, and the resulting q2 expression can be further maximized with respect to the energy E varying in the tunneling window. In one example, the energy leading to maximum q2 is
Moreover, the corresponding qM2 may be written as follows:
When neither the top nor the bottom 2D layers 106, 112 are degenerately doped, the tunneling window is given by Emin=ECT and Emax=EVB, in which case the EM defined in Equation (32) belongs to the tunneling window, and the maximum value of q2 is given by Equation (33). If either the top or the bottom 2D layer 106, 112 is degenerately doped, the Fermi levels may become the edges of the tunneling window, and the maximum value of q2 may be smaller than in Equation (33).
A considerable simplification in the evaluation of Equation (30) is obtained for qM2<<1/Lc2, in which case Equation (16) returns to SF(q)≈πLc2, so that by substituting SF(q) into Equation (29) and then into Equation (15), the expression for the current simplifies to:
where Emin=max {ECT,EFT} and Emax=min{EVB, EFB} define the tunneling window.
It should be understood that Equation (34) is consistent with a loss of momentum conservation, such that the current is simply proportional to the integral over the tunneling window of the product of the density of states in the top and bottom 2D layers 106, 112. Because the density of states is energy independent for a parabolic effective mass approximation, the current is proportional to the width of the tunneling window. In physical terms, Equation (34) corresponds to a situation where the scattering produces a complete momentum randomization during the tunneling process.
As long as the top 2D layer 106 is not degenerate, Emin=ECT and the tunneling window widens with the increase of the top gate voltage VTG. Hence, as represented in Equation (34), the current increases linearly with VTG. However, when the tunneling window increases to such an extent that qM2 becomes comparable to or larger than 1/Lc2, then part of the q values in the integration of Equation (30) may belong to the tail of the spectrum SF(q) defined in Equation (16). As a result, their contributions to the current become progressively diminished. In terms of the example Thin-TFET 100, while the tunneling window grows, the magnitude of the wave-vectors in the top and bottom 2D layers 106, 112 also increases, and, consequently, the scattering can no longer provide momentum randomization for all possible wave-vectors involved in the tunneling process. In such circumstances, the current first increases sub-linearly with VTG and eventually saturates for large-enough VTG values.
The 2D materials of the top and bottom 2D layers 106, 112 used in many of the examples herein are the hexagonal monolayer MoS2 and WTe2. The band structure for MoS2 and WTe2 may be determined using a density functional theory (DFT), which shows that these materials have a direct bandgap with the band edges for both the valence and the conduction band residing at the K point in the 2D Brillouin zone.
With respect now to
In some examples, the top gate 102 of the example Thin-TFET comprises Aluminum, which has a work function of 4.17 eV. Likewise, in some examples, the back gate 108 of the example Thin-TFET comprises p++ Silicon, which has a work function of 5.17 eV. Further, in some examples, the top and bottom oxide layers 104, 110 have an effective oxide thickness (EOT) of 1 nm. In one example, the top 2D layer 106 comprises hexagonal monolayer MoS2, while the bottom 2D layer 112 comprises hexagonal monolayer WTe2. For purposes of discussion here, and at least in some examples, an n-type and p-type doping density of 1012 cm−2 by impurities and full ionization are present in, respectively, the top and bottom 2D layers 106, 112, and the relative dielectric constant of the interlayer 118 material is 4.2 (e.g., boron nitride). In one example, the voltage VDS between the drain 114 and the source 116 is set to 0.3 V, and the back gate 108 is grounded unless stated otherwise. Further, the value of MB,0 can in some cases be determined from testing. In other cases, however, the value of MB,0 may be set to 0.01 eV, which is consistent with other applications, such as, for example, in a graphene/hBN system. For purposes of discussion herein, the value of MB,0 is set to 0.01 eV.
With respect to
In the linear region of this example, the tunnel current density IDS exhibits an approximately linear dependence on VTG and, indeed, the current is roughly proportional to the energy tunneling window, as discussed above and represented in Equation (34). This follows because the tunneling window is small enough that the condition qM2<<1/Lc2, is fulfilled. In this linear region, the tunnel current density IDS is proportional to the long-wavelength part of the scattering spectrum SF(q) (i.e., small q values). Hence the current may increase with the correlation length Lc, as expected based on Equation (34). The super-linear behavior of the tunneling current density IDS at small top gate voltage VTG values observed in
It should be understood that energy broadening and band tails have already been recognized as a fundamental limit to the SS of band-to-band tunneling transistors, and are not a specific concern of the example Thin-TFET 100. Further, as already mentioned above, the band tails in three-dimensional (3D) semiconductors have been investigated by using thermal measurements and are described in terms of the so called Urbach parameter E0. Values for the Urbach parameter E0 comparable to room temperature thermal energy (i.e., kBT≅26 meV) have been reported for GaAs and InP. By contrast, energy broadening and band tails in 2D materials play an important role in the minimum SS attainable by Thin-TFETs, and no data has been reported, synthesized, or utilized for band tails in monolayers of TMDs.
The example Thin-TFET 100 is a new steep slope transistor based on interlayer tunneling between two 2D semiconductor materials, namely, the top and bottom 2D layers 106, 112. The example Thin-TFET 100 allows for a very steep subthreshold region, and the SS may ultimately be limited by energy broadening in the two 2D materials comprising the top and bottom 2D layers 106, 112. The energy broadening can have different physical origins, such as, for example, disorder, charged impurities in the top and bottom 2D layers 106, 112 or in the surrounding materials, phonon scattering, and microscopic roughness at interfaces. Energy broadening has been accounted for here by assuming a Gaussian energy spectrum with no explicit reference to a specific physical mechanism. Moreover, while a possible rotational misalignment between the top and bottom 2D layers 106, 112 may affect the absolute value of the tunneling current, the misalignment does not significantly degrade the steep subthreshold slope offered by the example Thin-TFET 100, which may be the most crucial figure in terms of merit for a steep slope transistor.
Optimal operation of the example Thin-TFET 100 may require a good electrostatic control of the top gate voltage VTG on the band alignments in the material stack, as shown for example in
Those having ordinary skill in the art will appreciate that the above description of the example Thin-TFET 100 does not explicitly account for possible traps or defects assisted tunneling, which are known to be a serious hindrance to tunnel-FETs exhibiting a SS better than 60 mV/dec. Further, from a fundamental viewpoint, 2D crystals may offer advantages over their 3D counterparts because they are inherently free of broken/dangling bonds at the interfaces.
In short, the example Thin-TFET 100 is based on interlayer tunneling between two 2D materials. The Thin-TFET 100 has a very steep turn-on characteristic because the vertical stack of 2D materials having an energy gap is allows for the most effective, gate-controlled crossing and uncrossing between the edges of the bands involved in the tunneling process.
In view of the foregoing, various operating scenarios for the example Thin-TFET 100 were determined using an effective barrier height of the van der Waals gap between the top 2D layer 106 (in this example, SnSe2) and the bottom 2D layer 112 (in this example, WSe2) of 1.0 eV, a tunneling direction effective electron mass in van der Waals gap of m0, a tunneling distance of 0.3 nm, a correlation length Lc of scattering of 10 nm, an R.M.S. value of the scattering potential (i.e., matrix element) of 0.05 eV, an energy broadening of the density-of-state of 10 meV, and a top and bottom oxide EOT of 1 nm. For the purposes of brevity and avoiding redundancy, the results of such operating scenarios shown in
Based on these conditions and no contact resistance,
Based on the capacitance model 140 shown in
Still further, as shown in
With reference now to
The article by M. Li, et al., “Single particle transport in two-dimensional heterojunction interlayer tunneling field effect transistor,” J. of Applied Physics 115, 074508 (2014), is hereby incorporated by reference in its entirety. Further, although certain example methods and apparatus have been described herein, the scope of coverage of this patent is not limited thereto. On the contrary, this patent covers all methods, apparatus, and articles of manufacture fairly falling within the scope of the appended claims either literally or under the doctrine of equivalents.
This application is a non-provisional application claiming priority from U.S. Provisional Application Ser. No. 62/118,980, filed Feb. 20, 2015, entitled “Two-Dimensional Heterojunction Interlayer Tunneling Field Effect Transistors” and incorporated herein by reference in its entirety.
This invention was made with government support under Contract FA9550-12-1-0257 awarded by the Air Force Office of Scientific Research. The government has certain rights in the invention.
Number | Date | Country | |
---|---|---|---|
62118980 | Feb 2015 | US |