Sie sind auf Seite 1von 10

Superlattices and Microstructures 86 (2015) 211–220

Contents lists available at ScienceDirect

Superlattices and Microstructures


journal homepage: www.elsevier.com/locate/superlattices

A new analytical threshold voltage model of cylindrical gate


tunnel FET (CG-TFET)
S. Dash a, G.P. Mishra b,⇑
a
Dept. of Electronics & Communication Engg., Device Simulation Lab, Institute of Technical Education & Research, Siksha ‘O’ Anusandhan University,
Khandagiri, Bhubaneswar, India
b
Dept. of Electronics & Instrumentation Engg., Device Simulation Lab, Institute of Technical Education & Research, Siksha ‘O’ Anusandhan University,
Khandagiri, Bhubaneswar, India

a r t i c l e i n f o a b s t r a c t

Article history: The cylindrical gate tunnel FET (CG-TFET) is one of the potential candidates for future
Received 5 June 2015 nano-technology, as it exhibit greater scaling capability and low subthreshold swing (SS)
Received in revised form 17 July 2015 as compared to conventional MOSFET. In this paper, a new analytical approach is proposed
Accepted 20 July 2015
to extract the gate dependent threshold voltage for CG-TFET. The potential distribution and
Available online 21 July 2015
electric field distribution in the cylindrical channel has been obtained using the 2-D
Poisson’s equation which in turn computes the shortest tunneling distance and tunneling
Keywords:
current. The threshold voltage is extracted using peak transconductance change method
CG-TFET
Drain current
based on the saturation of tunneling barrier width. The impact of scaling of effective oxide
Shortest tunneling distance thickness, cylindrical pillar diameter and gate length on the threshold voltage has been
Transconductance investigated. The consistency of the proposed model is validated with the TCAD simulated
results. The present model can be a handful for the study of switching behavior of TFET.
Ó 2015 Elsevier Ltd. All rights reserved.

1. Introduction

Increased leakage current (IOFF) and subthreshold swing (SS) are the two major hindrances in the downscaling process of
complementary metal oxide semiconductor (CMOS) technology in the low-power applications [1–6]. Tunnel FET (TFET) is
one of the emerging alternative for conventional MOSFET beyond 90 nm regime due to its low IOFF and SS below
60 mV/decade. But on the contrary it results low drain current due to the tunneling of charge carriers at the source–channel
interface based on non-local Band-to-Band tunneling (BTBT) phenomena [7–10]. However a number of techniques such as
bandgap engineering, gate engineering, use of compound semiconductors, use of high-k dielectric have been employed to
improve the drain current in TFET structure [11–14]. Similarly different TFET models based on multi-gate structure have
been developed in recent times, to improve the device performance and scaling capability [15–17]. Cylindrical gate tunnel
FET (CG-TFET) provides enhanced electrostatic performance and optimum scaling capability due to its typical characteristics
length [18,19].
The investigation of device performance and the scaling optimization of CG-TFET are mostly focused using TCAD simu-
lations and experiments. But analytical models can give a brief insight for the design and fabrication of TFET devices. Barely a
few numbers of analytical models have been proposed to calculate surface potential, drain current of CG-TFET [20–23]. This

⇑ Corresponding author.
E-mail address: gurumishra@soauniversity.ac.in (G.P. Mishra).

http://dx.doi.org/10.1016/j.spmi.2015.07.049
0749-6036/Ó 2015 Elsevier Ltd. All rights reserved.
212 S. Dash, G.P. Mishra / Superlattices and Microstructures 86 (2015) 211–220

is due to the fact that, the physics related to the carrier tunneling of CG-TFET is completely different from standard
CG-MOSFETs.
The gate dependent threshold voltage is one of the key electrical parameters of CG-TFET, as it switches the device. In this
paper, an analytical gate threshold voltage model for CG-TFET is developed based on the saturation of the shortest tunneling
distance with respect to gate voltage. The threshold voltage is extracted using peak transconductance change (TC) method
and a methodical comparison with the constant current (CC) method is reported. Also the effect of the variation of effective
gate oxide thickness, cylindrical pillar diameter and gate length on the threshold voltage and drain current has been studied
extensively. The model accuracy is validated using TCAD Sentaurus simulator [24]. To study the transport behavior of
CG-TFET, non-local path BTBT model, Shockley–Read–Hall (SRH) Recombination model, band-gap narrowing concentration
models, and quantum potential model are considered.

2. Analytical model and its validation

2.1. Device structure & electrostatic analysis

Fig. 1 displays the structural architecture of n-channel Si based CG-TFET with 50 nm channel length (L) and 20 nm sour-
ce/drain length (LS/LD). The source and drain of the model is uniformly doped with trivalent and pentavalent impurity of the
order NS = ND = 1020 cm3 respectively. The gated channel region is very lightly doped with NC = 1015 cm3. The cylindrical
body diameter (tsi) and Equivalent Oxide Thickness (EOT) of SiO2 dielectric layer is considered as 10 nm and 2 nm respec-
tively. Here the work-function (Um) for the material used as metal gate contact is assumed as 4.2 eV for the model which
spreads over the 50 nm effective channel.
Threshold voltage is one of the significant electrical parameter of a device which influences the critical voltage values
characterizing IC digital circuits. In TFET, it signifies the transition between weak tunneling and strong tunneling of charge
carriers at the source–channel interface [25]. To develop the threshold voltage model of CG-TFET, we need to resolve the 2-D
Poisson’s equation for the entire channel region. The electrostatic distribution and electric field along the lateral direction of
N-TFET can be obtained by solving Poisson’s equation in the cylindrical coordinate system from the weak inversion
(OFF-state) to the onset of the strong inversion (ON-state). This further helps us to develop the analysis of drain current using
band-to-band generation rate and shortest tunneling distance.
The potential distribution u(r, z) in the cylindrical channel can be realized by 2-D Poisson’s equation with parabolic
approximation [1]. Here the effect of the mobile carriers and trapped charges in the oxide layer has been neglected.
 
1 @ @ @ 2 uðr; zÞ qNc t si
r uðr; zÞ þ ¼ ; 06r6 & 06z6L ð1Þ
r @r @r @z2 si 2

where si is the dielectric constant of silicon. The surface potential along the channel–gate dielectric interface can be
obtained using the boundary conditions defined for the cylindrical structure [1] as

@ 2 us ðzÞ ðV GS  V FB  us ðzÞÞ qN c
þ ¼ ð2Þ
@z2 2
lc si
Here V GS and V FB represents the gate-to-source voltage and flat-band voltage respectively. The characteristics length of
cylindrical gate structure (lc ) is defined as
vffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
 ffi
u
ut2 si ln 1 þ 2EOT
t si t si
lc ¼ ð3Þ
8ox
Eq. (2) can be solved using the boundary conditions at the source–channel and drain–channel interface.

Fig. 1. (a) Schematic cross-sectional and (b) side-view of n-type CG-TFET.


S. Dash, G.P. Mishra / Superlattices and Microstructures 86 (2015) 211–220 213

 
KT NS
us ðzÞz¼0 ¼  log ð4Þ
q ni
 
KT ND
us ðzÞz¼L ¼ V DS þ log ð5Þ
q ni

where V DS is the drain-to-source voltage, ox is the relative permittivity of SiO2, NS/ND is the doping profile of source/drain, KT
q
is the thermal voltage and ni is the intrinsic carrier concentration of silicon. For this model, the approximations are

KT
NS ¼ ND ¼ N ¼ 1020 cm3 & Vt ¼ ¼ 0:0259 V:
q
Solving Eq. (2), the surface potential in the lateral direction of cylindrical channel is found to be
!
2
z lz qNc lc
us ðzÞ ¼ Ae þ Be c  V FB  V GS þ
lc ð6Þ
si
where
" !   #
qNc lc   KT 
2
1 lL N lL
A¼   V DS þ V FB  V GS þ 1e c þ log 1þe c ð7Þ
2 sinh L si q ni
lc

" !   #
qN c lc  lL  KT 
2
1 N L
B¼   V DS þ V FB  V GS þ ec  1  log 1 þ e lc ð8Þ
2 sinh L si q ni
lc

Here the electrostatic potential distribution along the z-direction is modulated by gate bias as CG-TFET is a gated P-I-N
diode. The electric field along the lateral direction plays a significant role in the drain current analysis and thus in the thresh-
old voltage extraction. This is primarily because of the maximum BTBT charge carrier flow along the z-axis of the channel
[23]. The electric field distribution can be expressed within the range 0 6 z 6 L as
   
@ us A ðlz Þ B ðlz Þ
Ez ðr; zÞ ¼  ¼ ec þ e c ð9Þ
@z lc lc
Fig. 2a and b illustrate the potential and lateral electric field distribution of n-channel CG-TFET for different gate voltages.
The figure shows the improvement in surface potential with the increase in gate bias from 0.3 V to 0.5 V which is due to the
enhanced gate control on the intrinsic channel. Here the drain voltage and work function of material is kept constant. The
potential at the source/drain interface is dependent on built-in-potential which is controlled by the doping profile of sour-
ce/drain. However the slope of the surface potential curve increases due to high gate voltage at constant VDS. This change in
potential critically decides the shortest tunneling distance in the tunneling of charge carriers. The constant potential is
obtained at the middle of channel for a fixed VGS as gate control becomes saturated. This leads to zero lateral electric field
along the channel as shown in Fig. 2b. The model produces highest electric field at the tunneling junction (source–channel
interface) for high gate voltage which on the contrary reduces the shortest tunneling distance. But at the drain end, the low-k
dielectric SiO2 suppresses the ambipolar leakage conduction due to low electric field.

Fig. 2. (a) Surface potential distribution and (b) lateral electric field distribution of n-channel CG-TFET along the channel for different gate voltages.
214 S. Dash, G.P. Mishra / Superlattices and Microstructures 86 (2015) 211–220

Fig. 3. Energy band diagram of the proposed model in (a) OFF-state (VGS = 0 V) and (b) ON-state (VGS = 1 V).

2.2. Shortest tunneling distance analysis

The energy band diagram for the n-channel CG-TFET model of both OFF-state (VGS = 0 V) and ON-state (VGS = 1 V) has been
presented in Fig. 3(a) and (b). It is observed that the charge carriers tunnel through the source–channel interface in the
ON-state as the conduction band of channel and the valence band of source are in-line to each other [23]. This outline an
inflection point in the z-axis known as shortest tunneling distance (lt) which can be modulated by gate-to-source voltage.
 
 KT NS
ðus Þv al ðzÞ z¼0
¼ us ð0Þ ¼  log ð10Þ
q ni
!
i Eg lt lt qN l
2
Eg
uscond ðzÞ ¼ us ðlt Þ þ ¼ Aelc þ Belc  V FB  V GS þ c c þ ð11Þ
z¼lt q si q

where uscond ðzÞ and usv al ðzÞ denotes the potential of conduction band of channel and valence band of source for the n-channel
CG-TFET model. The shortest tunneling distance can be evaluated as
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi!
2
kþ k  4AB
lt ¼ lc loge ð12Þ
2A

where
2  
qN c lc Eg KT NS
k ¼ V FB  V GS þ þ  log ð13Þ
si q q ni

Fig. 4. Shortest tunneling distance variation w.r.t. gate voltages at constant drain voltage due to scaling of (a) effective oxide thickness and (b) Si pillar
diameter.
S. Dash, G.P. Mishra / Superlattices and Microstructures 86 (2015) 211–220 215

The shortest tunneling distance (lt) as a function of gate bias for different effective gate oxide thickness and silicon cylin-
drical pillar diameter are displayed in Fig. 4a and b. As the applied gate voltage increases past threshold value, the major
charge carrier tunnel through the source–channel interface and thus reduces the shortest tunneling distance due to superior
gate control. As the effective gate oxide thickness downscaled from 4 nm to 2 nm, the capacitive coupling effect between
gate and channel improves which curtails the shortest tunneling distance for a constant drain voltage. Similarly lt reduces
further with decrease in silicon body thickness from 14 nm to 10 nm, as the impact of gate voltage on channel increases
at low tsi. The reduction in the shortest tunneling distance at constant gate bias improves tunneling current due to larger
tunneling volume. Here the BTBT process is assumed to be extended up to the drain end along the channel and remain inde-
pendent of the variation of device parameters.

2.3. Tunneling current and transconductance

Here Kane’s nonlocal BTBT model [26] has been taken into consideration for the evaluation of tunneling current through
the channel for the interval of [z1 = lt, z2] and is expressed as [17]
Z t si Z z2 
2 Eg Ak Ez ðBEkgqzÞ
ID ¼ q e dz dr ð14Þ
t
 2si lt qz

where Ak and Bk are Kane’s tunneling parameters and the values taken in this model are 2.8  1019 m1/2 V5/2 s1 and
3.2  109 V/m respectively. Similarly Eg denotes the band-gap energy of silicon material and affect the shortest tunneling
distance significantly. Substituting electric field distribution along z-axis in the above equation, we get
2 0 1 0 1 3
Z z2
qB
ðl1  E k Þz Z z2 qB
ðl1 þ E k Þz
Ae c g Be c g
ID ¼ tsi Eg Ak 4 @ Adz þ @ Adz5 ð15Þ
lt lc z lt lc z

The drain current finally can be obtained by neglecting the effect of polynomial terms within the interval of [z1 = lt, z2] and
found to be
20 1 0 1 3
tsi Eg Ak 4@ A A B 
ID ¼ qBk
Pz2  Pz1  @qB A Qz  Qz 5 ð16Þ
lc E
 l1 E
k
þ l1 2 1
g c g c

where
qB qB
ð 1  E k Þz ðl1 þ E k Þz
e lc g e c g
Pz ¼ and Q z ¼ ð17Þ
z z
It has been observed that the coefficient P z2  Pz1 and Q z2  Q z1 . The simplified drain current is obtained by neglecting
the insignificant effects of Pz2 and Q z2 as
20 1 0 13
tsi Eg Ak 4@ AP z1 A @ BQ z1 A5
ID ¼ þ 1 qB ð18Þ
lc 1
l
 qB
E
k
l
þ Ek
c g c g

The transconductance of the model can be obtained by the 1st order differentiation of drain current w.r.t. gate voltage for
a constant drain voltage and is expressed as
 L   L 
dID Y 1 elc  1  Y 2 elc  1 dY 1 dY 2
gm ¼ ¼   þA þB ð19Þ
dV GS L
2 sinh lc dV GS dV GS

where

t si Eg Ak Pz1 tsi Eg Ak Q z1
Y1 ¼   & Y2 ¼   ð20Þ
1  lcEqBg k 1 þ lcEqBg k

Similarly the transconductance generation factor (TGF) is one of the important figure of merit of the device and is defined
as the ratio of transconductance to the drain current for a fixed value of gate voltage (gm/ID). TGF also gives an idea about the
gain of the device per unit power dissipation.
In Fig. 5, ID  VGS transfer characteristics has been shown in both linear and logarithmic scale for a constant drain voltage.
The effect of scaling of gate oxide thickness and silicon cylindrical pillar diameter on drain current performance has been
illustrated in Fig. 6. The current characteristics both in log scale and linear scale for constant VDS are displayed in the left
and right vertical axis respectively. It is clearly evident that the drain current increases with the reduction in the gate oxide
thickness and Si pillar diameter. This is due to the decrease in the shortest tunneling distance and improvement of non-local
216 S. Dash, G.P. Mishra / Superlattices and Microstructures 86 (2015) 211–220

Fig. 5. Drain current (ID) as a function of gate voltage (VGS) for the CG-TFET model.

Fig. 6. ID  VGS characteristics for various (a) effective gate oxide thickness and (b) silicon cylindrical pillar diameter at constant drain voltage.

Fig. 7. TGF as a function of gate voltage for different values of (a) effective gate oxide thickness and (b) silicon cylindrical pillar diameter at VDS = 1 V and
L = 50 nm.

BTBT volume. The nonlinear current curve in CG-TFET is influenced by the variation of EOT and tsi, which is due to the vari-
ation of gate control on intrinsic region at constant drain bias.
The scaling of effective gate oxide thickness and silicon cylindrical pillar diameter also results in improvement of
sub-threshold swing (SS). The average sub-threshold swing improves by 9 mV/dec i.e. from 50 mV/dec to 41 mV/dec with
S. Dash, G.P. Mishra / Superlattices and Microstructures 86 (2015) 211–220 217

downscaling of EOT from 4 nm to 2 nm. Similarly scaling of tsi results in decrease of SS by 5 mV/dec (from 46 mV/dec to
41 mV/dec). The above results are compared and validated with the existing experimental results [27].
Fig. 7 displays the variation of TGF w.r.t. gate voltage for various values of EOT and tsi for 50 nm gate length. It is worth
noting that TGF reduces with the increase in gate voltage as TGF is maximum at the weak tunneling region and reduces faster
as we approach strong tunneling of charge carriers. The above figure also illustrates the effect of downscaling of EOT and tsi
on the TGF performance. The small value of gate oxide thickness (EOT = 2 nm) and Si pillar diameter (tsi = 10 nm) gives the
lowest TGF at constant drain voltage which indicates that the tunneling of carriers at the interface starts for low VGS.

2.4. Threshold voltage extraction and model validation

The ID  VGS characteristics of CG-TFET displays the non-linear behavior which mainly controlled by narrowing the short-
est tunneling distance (lt) in the ON-state as shown in Figs. 5 and 6. The shortest tunneling distance primarily varies w.r.t.
applied gate voltage. However the tunneling of charge carriers in the ON-state need a minimum amount of drain voltage
irrespective of gate bias which reverse bias the P-I-N diode. Threshold voltage of a tunnel FET is the gate voltage for which
the shortest tunneling distance gradually starts to saturate with the applied gate bias. This is due to the saturation of nar-
rowing of energy barrier between valence band and conduction band.
The threshold voltage of conventional MOSFET can be computed using several approaches such as constant current
method (CC), linear extrapolation method (LE), split C-V method (SC), Fowler and Hartstein method (FH) and transconduc-
tance peak method (TC) [28]. In constant current (CC) method the threshold voltage is the value of the gate voltage for which
drain current IDcon ¼ 1013 A=lm [11,14]. But this method uses an optimized value of drain current which may not produce
accurate threshold voltages for TFET models and thus carry less practical meaning [29]. The ID  VGS characteristics of
CG-TFET provides the value of threshold (Vth = VGS) in a graphical manner at which drain current reaches a constant value
of 1013 A=lm. The threshold voltage curve is shown in Fig. 8 at a constant drain voltage which corresponds to shortest tun-
neling path of around 5–6 nm.
In this paper, the more accurate peak transconductance change (TC) method has been used to extract the threshold volt-
age of the n-channel CG-TFET. In this method, the threshold voltage of CG-TFET is the value of gate voltage which corre-
 
dg m
sponds to the maximum of the 1st order differentiation of transconductance dV GS
. This threshold voltage extraction
method detects the gate voltage where the shortest tunneling path gets saturated. Also it exhibits the transition between
strong and weak controls of the tunneling energy barrier width [25]. The change in transconductance can be analytically
obtained by differentiating Eq. (19) as
 L   L 
dg m T 1 elc  1  T 2 elc  1 dT 1 dT 2
¼   þA þB ð21Þ
dV GS L
sinh lc dV GS dV GS

where
2At si Eg Ak T 3 Pz1 2At si Eg Ak T 3 Q z1
T1 ¼  pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi and T 2 ¼   pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi ð22Þ
2 2
k þ k  4AB k þ k  4AB

Here constant T 3 is defined as

Fig. 8. Extraction of threshold voltage of CG-TFET by peak transconductance change method for (a) VGS = 1 V and (b) VGS = 1.5 V at constant drain voltage.
218 S. Dash, G.P. Mishra / Superlattices and Microstructures 86 (2015) 211–220

   pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi L

2
4AT 4 sinh L
lc
þ k þ k  4AB 1  elc
T3 ¼   ð23Þ
8A2 sinh lLc

where

2      3
L lL
L
6 k sinh lc
þ A 1  elc  B 1  e c
7
T 4 ¼ 1  4 pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi   5 ð24Þ
2
k  4AB sinh lLc

However the peak value of transconductance change can be realized by finding the inflection point of the curve. The point
is the byproduct of finding 1st order derivative of the result and equating to zero. The process has been illustrated in the
Fig. 8 and validated by the TCAD simulator.
Fig. 8 display the drain current (ID), transconductance (g m ) and change in transconductance (gm ) for the proposed model
at a maximum gate voltage of 1 V and 1.5 V respectively. The transconductance and change in transconductance are obtained
by the 1st order and 2nd order differentiation of drain current w.r.t. gate bias keeping drain voltage constant. Here the line
represents the results of analytical threshold model and are accurately validated by the symbols, signifies TCAD simulation.
Here the threshold voltage of CG-TFET is defined by the peak value of change in transconductance curve as shown in
Fig. 8a and b. The TC method of extracting threshold voltage is more precise with the simulation result and is independent
of gate length at constant drain voltage.
Fig. 9a displays the 2nd order differentiation of drain current w.r.t. gate voltage for 50 nm channel length with different
gate oxide thickness (EOT = 2 nm, 3 nm and 4 nm). The maximum value of dgm/dVGS curve indicates the threshold value of
the structure. It is clearly evident that the non-linear behavior of rate of change in transconductance curve is influenced by
the change in oxide thickness. The downscaling of EOT from 4 nm to 2 nm yields the reduction in shortest tunneling path and
thus decrease in threshold voltage as illustrated in Fig. 9b. However the threshold voltage extracted using proposed TC
method is compared with constant current method and produces results very similar to that of TCAD simulation result.
The threshold voltage versus gate length at constant drain and gate voltage is shown in Fig. 10. The threshold values for
different gate length are extracted using 1st order differentiation of transconductance and CC method. Here the threshold
voltage extracted by the proposed TC method is 0.91 V which corresponds to the narrowing of the shortest tunneling dis-
tance and is much higher than the threshold voltage defined by the constant current method (Vth = 0.52 V at
ID ¼ 1013 A=lm). The proposed results show good coincidence to that of TCAD simulation results as compared to CC
approach. However the constant nature of the curve is due to the fact that the lateral electric field is maximum at the
source–channel interface irrespective of gate length as shown in Fig. 2b. The threshold voltage of CG-TFET is independent
of the gate length for a positive drain voltage.
Fig. 11 illustrates the threshold voltage variation of n-channel CG-TFET as a function of Si pillar diameter and effective
gate oxide thickness at VGS = 1.5 V and VDS = 1 V. When the Si pillar diameter is scaled down from 20 nm to 10 nm at a fixed
gate length, the threshold voltage reduces linearly as the narrowing of shortest tunneling distance saturates. However the
proposed method of extraction of the threshold voltage resembles with the simulated result as compared to CC method.
Similarly the threshold voltage variation in CG-TFET is sensitive to the gate oxide thickness. Low Vth is obtained for
EOT = 2 nm, because of the better capacitive control of the barrier width at the source–channel interface.

Fig. 9. (a) Threshold voltage extraction using TC method and (b) variation of threshold voltage for different values of effective gate oxide thickness with
VGS = 1 V and VDS = 1 V.
S. Dash, G.P. Mishra / Superlattices and Microstructures 86 (2015) 211–220 219

Fig. 10. Variation of threshold voltage due to change in channel length with VGS = 1 V and VDS = 1 V.

Fig. 11. Variation of threshold voltage w.r.t. (a) Si pillar diameter and (b) gate oxide thickness with L = 50 nm, VGS = 1.5 V and VDS = 1 V.

3. Conclusions

In this paper, the analytical expression for the shortest tunneling distance and drain current have been derived using the
solution of 2-D Poisson’s equation in the intrinsic channel. The gate threshold voltage is further extracted using proposed
transconductance change method based on the saturation of shortest tunneling distance. It is clearly evident that the pro-
posed method predicts the nature of switching behavior more precisely as compared to constant current method. However
the analytical model is also able to investigate the drain current and threshold voltage w.r.t. the variation in the gate length,
effective oxide thickness and silicon pillar diameter. Finally the proposed threshold voltage model for CG-TFET can provide
an insight for the future TFET technology in the low power applications.

References

[1] A. Kranti, S. Haldar, R.S. Gupta, Analytical model for threshold voltage and I–V characteristics of fully depleted short channel cylindrical/surrounding
gate MOSFET, Microelectron. Eng. 56 (2001) 241–259.
[2] K.K. Young, Short-channel effect in fully-depleted SOI MOSFETs, IEEE Trans. Electron. Dev. 36 (1989) 399–402.
[3] A.O. Adan, T. Naka, A. Kagisawa, H. Shimizu, SOI as a mainstream IC technology, in: Proc. of IEEE International SOI Conference, 1998, pp. 9–12.
[4] C.P. Auth, J.D. Plummer, Scaling theory for cylindrical fully-depleted, surrounding gate MOSFET’s, IEEE Electron Dev. Lett. 18 (1997) 74–76.
[5] J.P. Colinge, FinFETs and Other Multi-Gate Transistors, Springer, 2008.
[6] J.P. Colinge, Multiple-gate SOI MOFETs, Solid-State Electron. 48 (2004) 897–905.
[7] W.Y. Choi, B.G. Park, J.D. Lee, T.J.K. Liu, Tunneling field-effect transistors (TFETs) with subthreshold swing (SS) less than 60 mV/dec, IEEE Electron Dev.
Lett. 28 (2007) 743–745.
[8] L. Lattanzio, A. Biswas, L.D. Michielis, A.M. Ionescu, Abrupt switch based on internally combined band-to-band and barrier tunneling mechanisms,
Solid-State Electron. 65–66 (2011) 234–239.
[9] Q. Zhang, W. Zhao, S. Alan, Low-subthreshold-swing tunnel transistors, IEEE Electron Dev. Lett. 27 (2006) 297–300.
220 S. Dash, G.P. Mishra / Superlattices and Microstructures 86 (2015) 211–220

[10] Th. Nirschl, St. Henzler, J. Fischer, M. Fulde, A. Bargagli-Stoffi, M. Sterkel, Scaling properties of the tunneling field effect transistor (TFET): device and
circuit, Solid-State Electron. 50 (2006) 44–51.
[11] K. Boucart, A.M. Ionescu, Double-gate tunnel FET with high-j gate dielectric, IEEE Trans. Electron Dev. 54 (7) (2007) 1725–1733.
[12] Y. Wu, H. Hasegawa, K. Kakushima, K. Ohmori, T. Watanabe, A. Nishiyama, et al, A novel hetero-junction Tunnel-FET using Semiconducting silicide–
Silicon contact and its scalability, Microelectron. Reliab. 54 (5) (2014) 899–904.
[13] C.H. Shih, N.D. Chien, Design and modeling of line-tunneling field-effect transistors using low-bandgap semiconductors, IEEE Trans. Electron Dev. 61
(2014) 1907–1913.
[14] K.K. Bhuwalka, J. Schulze, I. Eisele, Scaling the vertical tunnel FET with tunnel bandgap modulation and gate work-function engineering, IEEE Trans.
Electron Dev. 52 (2005) 909–917.
[15] A. Pan, S. Chen, C.O. Chui, Electrostatic modeling and insights regarding multigate lateral tunneling transistors, IEEE Trans. Electron Dev. 60 (2013)
2712–2720.
[16] A.S. Verhulst, B. SoreÌ’e, D. Leonelli, W.G. Vandenberghe, G. Groeseneken, Modeling the single-gate, double-gate, and gate-all-around tunnel field-effect
transistor, J. Appl. Phys. 107 (2010) 024518-1–024518-8.
[17] M. Gholizadeh, S.E. Hosseini, A 2-D analytical model for double-gate tunnel FETs, IEEE Trans. Electron Dev. 61 (2014) 1494–1500.
[18] R. Narang, M. Saxena, R.S. Gupta, M. Gupta, Drain current model for a gate all around (GAA) p–n–p–n tunnel FET, Microelectron. J. 44 (2013) 479–488.
[19] R. Vishnoi, M.J. Kumar, Compact analytical drain current model of gate-all-around nanowire tunneling FET, IEEE Trans. Electron Dev. 61 (2014) 2599–
2603.
[20] M. Yadav, A. Bulusu, S. Dasgupta, Two dimensional analytical modeling for asymmetric 3T and 4T double gate tunnel FET in sub-threshold region:
potential and electric field, Microelectron. J. 44 (2013) 1251–1259.
[21] M.J. Lee, W. Choi, Analytical model of single-gate silicon-on-insulator (SOI) tunneling field-effect transistors (TFETs), Solid-State Electron. 63 (2011)
110–114.
[22] B. Bhusan, K. Nayak, DC compact model for SOI tunnel field-effect transistors, IEEE Trans. Electron Dev. 59 (2012) 2635–2642.
[23] S. Dash, G.P. Mishra, A 2-D analytical cylindrical gate tunnel FET (CG-TFET) model: impact of shortest tunneling distance, Adv. Natl. Sci.: Nanosci.
Nanotechnol. 6 (2015) 035005-1–035005-10.
[24] Sentaurus Device User Guide, Synopsys, Inc., Mountain View, USA, 2010.
[25] K. Boucart, A.M. Ionescu, A new definition of threshold voltage in tunnel FETs, Solid-State Electron. 52 (2008) 1318–1323.
[26] E.O. Kane, Zener tunneling in semiconductors, J. Phys. Chem. Solids 12 (2) (1960) 181–188.
[27] R. Gandhi, Z. Chen, N. Singh, K. Banerjee, S. Lee, CMOS-compatible vertical-silicon-nanowire gate-all-around p-type tunneling FETs with 650-mV/
decade subthreshold swing, IEEE Electron Dev. Lett. 32 (2011) 1504–1506.
[28] R. Booth, M. White, H. Wong, T. Krutsick, The effect of channel implants on MOS transistor characterization, IEEE Trans. Electron Dev. 34 (1987) 2501–
2509.
[29] L.Y. Chen, Z.H. Ming, Z.Y. Ming, H.H. Yong, W. Bin, L.Y. Le, Z.C. Yu, Two-dimensional threshold voltage model of a nanoscale silicon-on-insulator
tunneling field-effect transistor, Chin. Phys. B 22 (2013) 038501-1–038501-6.

Das könnte Ihnen auch gefallen