Uploaded by Rajesh Pathak

Msfet for low power requirements

advertisement
Journal of ELECTRONIC MATERIALS, Vol. 48, No. 10, 2019
https://doi.org/10.1007/s11664-019-07483-1
Ó 2019 The Minerals, Metals & Materials Society
Effect of Ferroelectric Thickness Variation in Undoped
HfO2-Based Negative-Capacitance Field-Effect Transistor
BHASKAR AWADHIYA ,1,2 PRAVIN N. KONDEKAR,1,3
and ASHVINEE DEO MESHRAM1,4
1.—Nano-Electronics and VLSI Lab, Electronics and Communication Engineering Department,
Indian Institute of Information Technology, Design and Manufacturing, Jabalpur 482005, India.
2.—e-mail:
bhaskar.a@iiitdmj.ac.in.
3.—e-mail:
pnkondekar@iiitdmj.ac.in.
4.—e-mail:
1612304@iiitdmj.ac.in
Negative-capacitance field-effect transistors (NCFETs) are emerging devices
which have shown huge potential to replace classical field-effect transistors
because of their steep switching characteristic enabled by a ferroelectric stack.
The negative capacitance in ferroelectrics results in a voltage step-up action
which curtails the subthreshold swing below 60 mV/dec. The ferroelectric
thickness is a key design parameter which governs the operation of such devices and resulting circuits. We examine herein for the first time the effect of
the ferroelectric thickness of undoped HfO2-based negative-capacitance fieldeffect transistors on the device and circuit performance. Increasing the ferroelectric thickness yields higher gain but with increased probability of hysteresis. Also, depending upon the properties of the underlying transistor, at
low overdrive voltage, increase in the ferroelectric thickness beyond a certain
value may introduce a loss of saturation (negative differential resistance) in
the drain characteristic of the NCFET. Also, we design an NCFET-based
resistive load inverter and study the effect of thickness variation on the circuit
performance. The results of the analysis show that increasing the thickness
within a permissible limit increases the noise margin and reduces the power
dissipation of the designed circuit.
Key words: Negative-capacitance field-effect transistor (NCFET), negative
differential resistance (NDR), ferroelectric–dielectric (FE–DE),
subthreshold swing (SS)
INTRODUCTION
Scaling in conventional transistors has helped
improve their switching speed, packing density, and
cost. Despite all these advantages, it has instigated
problems related to short-channel effects, high
power dissipation, and increased leakage current
of such device, deteriorating the ION =IOFF ratio of
transistors. As the gate length of a transistor is
reduced, the supply voltage ðVDD Þ and threshold
voltage ðVTH Þ must be decreased proportionally to
keep the overdrive ðVDD VTH Þ high. As a
(Received January 19, 2019; accepted July 24, 2019;
published online August 6, 2019)
6762
consequence, the leakage current ðIOFF Þ of the
transistor starts to rise exponentially. Thus, to
maintain the balance between overdrive and leakage current, the subthreshold swing of the transistor needs to be lowered. However, thermodynamics
dictates that the subthreshold swing of a conventional metal–oxide–semiconductor field-effect transistor (MOSFET) is restricted by Boltzmann
tyranny,1 which means that it is not possible to
curtail the subthreshold swing below 60 mV/dec.
This limitation in the subthreshold swing occurs
because of current transport phenomena, which
involves thermionic injection of electrons2–4 over the
barrier. The subthreshold swing of MOSFETs cannot be scaled below 60 mV/dec, even if Cox tends to
1. There are two approaches to reduce the
Effect of Ferroelectric Thickness Variation in Undoped HfO2-Based Negative-Capacitance
Field-Effect Transistor
subthreshold swing. The first is to modify the
transport factor, i.e., the current transport phenomena in the channel. The tunnel field-effect transistor
(TFET)5,6 and green transistors7 use the band-toband tunneling8 approach to reduce the subthreshold swing below 60 mV/dec, whereas the impactionization MOS technique9 uses impact ionization to
achieve the same. However, the biggest disadvantage with tunnel field-effect transistors is that they
suffer from the problem of low ON-current and
ambipolarity, whereas the operating voltage for
IMOS technology is very high. The second approach
is to modify the body factor, i.e., reduce it to less
than 1. This can be achieved by using nanoelectromechanical (NEM) relay devices10–12 or negativecapacitance field-effect transistors.13–15 In NEM
relay devices, the point of instability between a
mechanical and electrical force is utilized to obtain a
supersteep subthreshold swing. However, these
devices suffer from limitations in terms of voltage
scaling and reliability. Negative-capacitance fieldeffect transistors utilize the property of negative
capacitance16–21 to modify the body factor and hence
achieve (SS) subthreshold swing below 60 mV/dec.
An NCFET can be thought of as a baseline MOSFET
having an external amplifier at the input of the gate
terminal, which means that it does not alter the
transport physics in the MOSFET and hence can
benefit from all future material research aimed at
improving the channel transport. It should be kept
in mind that the ferroelectric thickness22 and
temperature23 are important design parameters
for NCFETs. Increasing the ferroelectric thickness
results in a decrease of the ferroelectric capacitance,
which in turn increases the amplification factor of
the NCFET. However, it should be kept in mind
that the NCFET becomes prone to negative differential resistance at higher ferroelectric thickness
and lower gate bias, as discussed below.
BASICS OF NCFET
An NCFET can be considered as a classical FET
with a passive gate voltage amplifier. It includes an
additional ferroelectric oxide layer above the dielectric oxide, as shown in Fig. 1a. The ferroelectric
oxide layer is considered as a capacitor connected in
series with a classical FET. The ferroelectric capacitor is modeled as a voltage-dependent source,
which is a function of the gate charge. All the
simulation results presented herein are obtained by
incorporating the parameters for undoped HfO2 into
an existing model, available at nanoHUB. Amalgamating the BSIM424 model for a classical MOSFET
and the Landau–Khalatnikov theory of negative
capacitance25 in ferroelectrics results in the NCFET
model. The dynamics of the ferroelectric can be
explained by the Landau–Khalatnikov equation,26,27 which can be expressed as a charge versus
voltage equation as shown in Eq. 1:
6763
Fig. 1. (a) Schematic of 32-nm NCFET. (b) Equivalent capacitor
model of NCFET.
VF ¼ aQF þ bQ3F þ cQ5F þ q
dQF
;
dt
ð1Þ
where q is a constant which depends on the
ferroelectric material and the amplitude of the
applied voltage16 and a, b, and c are anisotropy
constants for ferroelectric materials. A schematic
and capacitor model of a 32-nm NCFET are shown
in Fig. 1a and b, respectively. From the capacitance
model, an expression for VMOS (under steady state
and when the ferroelectric is operating in the
negative-capacitance region) can be written as
VMOS ¼
AG ¼
VG
ð2Þ
VMOS ¼ AG VG þ AD VD ;
ð3Þ
jCF j
;
jCF j CMOS
þ jC
VD
;
1
CMOS
jCF j
F jCMOS
AD ¼
CD
CD
;
jCF j CMOS
ð4Þ
where AG and AD are the amplification factor and
drain coupling factor, respectively. For nonhysteretic operation of an NCFET, jCF j CMOS must
6764
Awadhiya, Kondekar, and Meshram
be greater than zero across the full span of operating bias.28 This leads to the requirements that
AG > 0 and AD < 0. It is worth noting that an increase
in the ferroelectric thickness increases the total
capacitance of the MOSFET by a factor of AG .
Hence, the total capacitance of an NCFET is greater
than that of a conventional MOSFET, which may
increase the delay if the damping constant is not
properly optimized.29 Also, an increase in the ferroelectric thickness beyond a critical value introduces
hysteresis30 and sometimes nonvolatility.31 Hence,
it is advisable to maintain the thickness of the
ferroelectric oxide below this critical thickness.32
The ferroelectric thickness at which jCF j becomes
equal to CMOS is termed the critical thickness.
Increasing tFE beyond tC introduces hysteresis into
the device characteristics, where
2
Q0
:
tC ¼ pffiffiffi
3 3 EC CMOS
ð5Þ
The critical thickness ðtC Þ is dependent on the
total gate capacitance ðCMOS Þ associated with the
underlying transistor and the properties of the
ferroelectric material. Recent studies have shown
ferroelectric behavior in doped HfO2,33 but here we
consider undoped HfO2 as the ferroelectric material.
Undoped HfO2, when deposited using physical
vapor deposition (PVD), shows dielectric behavior,
but when deposited using chemical vapor deposition
(CVD) it shows ferroelectric behavior; using
undoped HfO2 makes the deposition process easier
and hence simplifies the process control. The remanent polarization ðPR Þ and coercive field ðEC Þ are
considered to be 10 lC/cm2 and 1000 kV/cm, respectively,34,35 while the value of the anisotropy constant is obtained using the following relation
mentioned in Ref. 36:
pffiffiffi
pffiffiffi
3 3 VC
3 3 VC
and b ¼
;
ð6Þ
a¼
2 QO
2 Q3O
where VC ¼ EC tFE is the coercive voltage and QO ¼
PR A is the remanent charge.
DEVICE ANALYSIS
To explain the operation of NCFET-based circuits,
the device characteristics and their dependence on
the ferroelectric thickness must be discussed. We
study herein the VMOS –VG characteristics of
NCFETs with different ferroelectric thicknesses,
as shown in Fig. 2a. As the ferroelectric thickness is
increased, VMOS becomes greater than VG due to the
negative capacitance in the ferroelectric oxide. After
exceeding the critical thickness, i.e., tFE ¼ 10 nm,
hysteresis can be observed in the VMOS versus VG
characteristic of the NCFET. Figure 2b shows the
VFE –VG characteristics of NCFETs with different
ferroelectric thicknesses, where VFE ¼ VG VMOS .
As the thickness of the ferroelectric is increased,
VFE starts decreasing and goes negative. After
exceeding the critical thickness, hysteresis can be
observed in the VFE –VG characteristic of the
NCFET. Figure 2c shows the AG –VG characteristics
of the NCFET, where AG ¼ @VMOS =@VG . Increasing
the ferroelectric thickness increases the amplification factor up to 10 nm, after which hysteresis
starts to appear in the internal voltage amplification characteristics. Figure 2d and e show the ID –
VG and log ID –VG characteristics for the NCFET,
respectively. It is clearly visible that, as tFE
increases, the ON-current increases and the switching characteristic becomes steeper, because of the
voltage amplification contributed by the ferroelectric oxide layer. Increase in the ON-current
enhances the ION =IOFF for the NCFET, as presented
explicitly in Table I. Figure 2f shows the subthreshold swing versus gate voltage characteristic of the
NCFET. It is apparent that increasing the ferroelectric thickness (up to the critical thickness)
reduces the SS of the device. Furthermore, the
expression for the subthreshold swing37 of the
NCFET is shown below, obtained using an equivalent capacitor model of the NCFET:
Cdep
60 mV
1
1þ
:
ð7Þ
SSNCFET ¼
dec
AG
Cox
The results presented above clearly indicate the
occurrence of hysteresis at tFE ¼ 15 nm, hence only
results for ferroelectric thicknesses up to 10 nm are
discussed below.
Figure 3a shows the g0m –VG characteristics for the
NCFET. The transconductance is defined as the
variation in the drain current with respect to a
variation in the gate voltage. Increasing the ferroelectric thickness leads to an increase in the amplification factor and hence the transconductance of
the NCFET. An expression for the transconductance
of the NCFET is shown in Eq. 8, clearly showing
that the transconductance of the NCFET is greater
than that of a MOSFET by a factor of AG .
g0m ¼
@ID
@ID
@VMOS
¼
¼ gm AG ;
@VG @VMOS
@VG
ð8Þ
where gm is the transconductance of the internal
MOSFET, g0m is the transconductance of the
NCFET, and AG is the amplification factor of the
NCFET. Figure 3b shows the g0m2 –VG characteristics for the NCFET. The second-order transconductance is defined as the second-order variation in the
drain current with respect to a variation in the gate
voltage. The second-order transconductance of the
NCFET is A2G times greater than that of the
MOSFET, as seen from Eq. 9.
@ 2 ID
@ 2 ID
@VMOS 2
g0m2 ¼
¼
¼ gm2 A2G ; ð9Þ
2
@VG
@VG2
@VMOS
Effect of Ferroelectric Thickness Variation in Undoped HfO2-Based Negative-Capacitance
Field-Effect Transistor
6765
Fig. 2. (a) Variation of VMOS with VG for different values of tFE . (b) Variation of VFE with VG for different values of tFE . (c) Variation in internal
voltage amplification ð@VMOS =@VG Þ with VG for different values of tFE . (d) ID –VG characteristics on linear scale for different values of tFE . (e) ID –VG
characteristics on logarithmic scale for different values of tFE . (f) Subthreshold swing (SS) as a function of gate voltage ðVG Þ for different values of
tFE .
Table I. Threshold voltage and ON/OFF-current
ratio with thickness variation in NCFET
NCFET
tFE
tFE
tFE
tFE
=
=
=
=
0 nm
5 nm
10 nm
15 nm
VTH (V)
ION/IOFF
0.27
0.26
0.25
Hysteresis
5640.4
11063
17642
Hysteresis
where gm2 is the second-order transconductance of
the internal MOSFET, g0m2 is the second-order
transconductance of the NCFET. The gate voltage
at which the peak value of the second-order
transconductance is obtained is the threshold voltage.38 The threshold voltage of the NCFET with
different ferroelectric thicknesses is presented
explicitly in Table I. Increasing the ferroelectric
thickness leads to a leftwards shift of the peak value
of the transconductance at lower gate bias,
6766
Awadhiya, Kondekar, and Meshram
in VD , as seen in Fig. 5a. Hence, the current
reduces with a positive incremental change in
VD , thus the output characteristic exhibits
negative differential resistance (NDR), primarily originating because of the dependence
of the drain bias ðVD Þ with internal node bias
ðVMOS Þ through the drain capacitance ðCD Þ.
2.2.
For VOV ¼ 0:4 V, AG VG is the dominant
factor and nullifies the effect of AD VD ,
hence an increase in VD leads to a small
decrement in VMOS , as seen in Fig. 5b. Hence,
NDR is not observed and the current saturation phenomenon is observed in the drain
characteristics of the NCFET.
2.3.
For VOV ¼ 0:9 V, AG VG becomes dominant
and hence VMOS shows an even weaker
dependence on VD , as seen in Fig. 5c, hence
VMOS becomes large enough to impact VDSAT
significantly. In general, VDSAT can be expressed as ðVMOS VTH Þa where a is an
exponent that depends on the nature of the
current saturation mechanism in the transistor,39 lying between 1 and 2. For larger values
of a, as in the MOSFET case, it is observed
that, for VOV ¼ 0:9 V, the current saturation
phenomenon is observed. However, for even
large values of VOV , VDSAT will exceed VDD ,
hence ID will fail to saturate.
0
0
Fig. 3. (a) gm
–VG characteristics for different values of tFE . (b) gm2
–
VG characteristics for different values of tFE .
indicating a reduction in the threshold voltage with
increasing ferroelectric thickness. However, when
exceeding the critical thickness, hysteresis appears
in the transfer characteristics, hence the threshold
voltage estimation, in this case, is trivial.
Figure 4a, b, and c show the ID –VD output characteristic of the NCFET at VOV ¼ 0:1 V, 0.4 V, and
0.9 V, respectively. Further discussion on the drain
characteristics of the NCFET is presented below:
1.
2.
At tFE ¼ 5 nm, jCF j dominates over CD , hence
the amplification factor ðAG Þ will be dominant
over the drain coupling factor ðAD Þ, so the
internal node voltage is more dependent on
the gate bias, thus the ID –VD behavior of the
NCFET at lower ferroelectric thickness is similar to that of the baseline FET tFE ¼ 0 nm with
larger ON-current, which is due to the voltage
amplification effect.
At tFE ¼ 10 nm; jCF j decreases and becomes
comparable to CD , hence the amplification factor is comparable to the drain coupling factor, so
the relative values of VD and VG control VMOS :
2.1.
For VOV ¼ 0:1 V, AD VD dominates over
AG VG , hence VMOS decreases with increase
Figure 4d, e, and f show the g0ds –VD characteristics of the NCFET at VOV ¼ 0:1 V, 0.4 V, and 0.9 V,
respectively. Further discussion on the g0ds –VD
characteristics of the NCFET is presented below:
3.
4.
At tFE ¼ 5 nm, g0ds decreases with increase in
VD , therefore g0ds attains a small and positive
value at higher value of VD , which is because of
the saturation of the drain current with respect
to VD .
At tFE ¼ 10 nm, the behavior of the g0ds –VD
characteristic is strongly dependent on the
overdrive voltage in the NCFET.
4.1.
For VOV ¼ 0:1 V, negative output conductance
can be observed, which is caused by the
negative differential resistance40 in the drain
characteristics of the NCFET. A negative
value of g0ds can also be described by a large
negative value of the drain coupling factor
ðAD Þ, as shown in Fig. 5d.
4.2.
For VOV ¼ 0:4 V, g0ds decreases with an increase in the drain voltage. Furthermore, at
higher value of VD , g0ds saturates, mainly
because of the saturation in the drain characteristics of the NCFET. Also, the drain
coupling factor ðAD Þ has a small negative
value, as seen in Fig. 5e.
Effect of Ferroelectric Thickness Variation in Undoped HfO2-Based Negative-Capacitance
Field-Effect Transistor
6767
Fig. 4. (a) ID –VD characteristics with VOV ¼ 0:1 V for different values of tFE . (b) ID –VD characteristics with VOV ¼ 0:4 V for different values of tFE .
0
0
–VD characteristics with VOV ¼ 0:1 V for different values of tFE . (e) gds
–
(c) ID –VD characteristics with VOV ¼ 0:9 V for different values of tFE . (d) gds
0
–VD characteristics with VOV ¼ 0:9V for different values of tFE .
VD characteristics with VOV ¼ 0:4 V for different values of tFE . (f) gds
4.3.
For VOV ¼ 0:9 V, g0ds decreases with increase
in VD . Furthermore, at VD ¼ 0:9 V, g0ds saturates, mainly because of the saturation in the
drain characteristics. However, further increase in VOV may lead to nonsaturation in
the g0ds –VD characteristic and g0ds decreases
monotonically with an increase in VD . Also,
the drain coupling factor ðAD Þ has an even
smaller negative value, as seen in Fig. 5f.
ANALYSIS OF NCFET-BASED RESISTIVE
LOAD INVERTER
Figure 6a and b show a schematic and the
transfer characteristic of a resistive load inverter
with RD ¼ 10k X. The voltage amplification ðAG Þ
provided by the ferroelectric oxide enables steep
switching in the voltage transfer characteristic of
the resistive load inverter. Also, at tFE ¼ 5 nm and
10 nm, the gain of the NCFET increases, which in
6768
Awadhiya, Kondekar, and Meshram
Fig. 5. (a) VMOS –VD characteristics with VOV ¼ 0:1 V for different values of tFE . (b) VMOS –VD characteristics with VOV ¼ 0:4 V for different values
of tFE . (c) VMOS –VD characteristics with VOV ¼ 0:9 V for different values of tFE . (d) AD –VD characteristics with VOV ¼ 0:1 V for different values of
tFE . (e) AD –VD characteristics with VOV ¼ 0:4 V for different values of tFE . (f) AD –VD characteristics with VOV ¼ 0:9 V for different values of tFE .
turn increases the noise margin and hence provides
better noise immunity to the circuit. The noise
margin of a circuit is defined as its ability to endure
noise without triggering fictitious changes in the
output voltage. It is a quantitative measure of the
noise immunity of the circuit. The noise margin can
be computed from the dVout =dVin versus Vin characteristic of the resistive load inverter, which is
shown in Fig. 6c. The applied input biases at which
the slope of the voltage transfer characteristic is
equal to 1 ðdVout =dVin ¼ 1Þ are VIH and VIL . Also,
the value of the output at these points is VOL and
VOH , respectively. The obtained values of
VIH ; VIL ; VOH ; VIL ; NMH , and NML for tFE ¼ 0:5 and
10 nm are presented in Table II.
NMH ¼ VOH VIH ;
ð10Þ
NML ¼ VIL VOL ;
ð11Þ
where NMH is the noise margin high and NML is the
noise margin low. At tFE ¼ 5 nm and 10 nm, the
Effect of Ferroelectric Thickness Variation in Undoped HfO2-Based Negative-Capacitance
Field-Effect Transistor
6769
Fig. 6. (a) Schematic of NCFET-based inverter. (b) Voltage transfer characteristics of NCFET-based inverter. (c) Voltage gain ð@Vout =@Vin Þ
versus Vin characteristics of NCFET-based inverter. (d) DC power consumption of NCFET-based inverter.
Table II. Noise margin for resistive load inverter
Parameter
VIH (V)
VIL (V)
VOH (V)
VOL (V)
NMH (V)
NML (V)
42.4% less compared with the MOSFET-based
resistive load inverter ðtFE ¼ 0 nm).
tFE = 0 nm
tFE = 5 nm
tFE = 10 nm
CONCLUSIONS
0.4073
0.1215
0.8531
0.0671
0.4458
0.0543
0.2956
0.0984
0.8760
0.0433
0.5804
0.0551
0.1858
0.0894
0.8750
0.0337
0.6892
0.0557
We explore the effect of FE thickness variation on
the characteristics of undoped HfO2-based NCFET
devices and resulting circuits. The results clearly
indicate that an increase in the ferroelectric thickness may lead to hysteresis in the transfer characteristic, which limits increase of the thickness
beyond a critical value. Also, an increase in the
thickness introduces NDR into the output characteristics of the NCFET at low overdrive voltage,
which is responsible for a negative value of the
output conductance in the g0ds –VD characteristics.
The VMOS –VD and AD –VD characteristics of the
NCFET are further discussed to provide understanding of the erratic behavior of the drain characteristics of the NCFET at different overdrive
voltages. A resistive load inverter is designed using
the NCFET, and its noise margin and power
dissipation evaluated. It is observed that the
NCFET-based circuit shows high immunity to noise,
hence having a higher noise margin. Also, the
noise margin high and noise margin low for the
NCFET-based resistive load inverter have higher
values compared with the NCFET-based resistive
load inverter with tFE ¼ 0 nm. It is also observed
that the resistive load inverter using the NCFET
shows decreased leakage power. The results shown
in Fig. 6d indicate a decrease in the leakage power
for the NCFET-based resistive load inverter. For
VIN ¼ 0 V, the NCFET-based resistive load inverters
with tFE ¼ 5 nm and 10 nm exhibit leakage of
0.2247 lW and 0.1708 lW, respectively. Also the
power dissipation for the NCFET-based resistive
load inverter at tFE ¼ 5 nm and 10 nm is 24.6% and
6770
Awadhiya, Kondekar, and Meshram
NCFET-based circuit consumes less power compared with the conventional MOSFET circuit.
ACKNOWLEDGMENTS
20.
The authors would like to thank Dr. Muhammad
Abdul Wahab (Intel Corporation) and Prof.
Muhammad Ashraful Alam (Purdue University) for
providing the NCFET model available at nanoHUB.
21.
23.
REFERENCES
24.
1. V.V. Zhirnov and R.K. Cavin, Nat. Nanotechnol. 3, 2 (2018).
2. M.S. Lundstrom, in 2006 IEEE International SOI Conference Proceedings, pp. 1–3 (2006).
3. S.M. Sze, Physics of Semiconductor Devices, 1st ed. (New
York: Wiley, 1969).
4. A.M. Ionescu and H. Riel, Nature 479, 7373 (2011).
5. K. Boucart and A.M. Ionescu, IEEE Trans. Electron Devices.
54, 7 (2007).
6. A.C. Seabaugh and Q. Zhang, in Proc. IEEE, pp. 2095–2110
(2010).
7. K. Jeon, W.Y. P. Loh, C.Y. Patel, J. Kang, A. Oh, C.
Bowonder, C.S. Park, C. Park, P. Smith, H.H. Majhi, R.
Tseng, T.S.K. Jammy, T. Liu and C. Hu, in 2010 Symposium
on VLSI Technology, pp. 121–122 (2010).
8. C. Zener, in Proceedings of Royal Society: A Mathematical,
Physical, and Engineering Sciences, pp. 523–529 (1935).
9. K. Gopalakrishnan, P.B. Griffin, and J.D. Plummer, in Digest
International Electron Devices Meeting, pp. 289–292 (2002).
10. N. Abele, R. Fritschi, K. Boucart, F. Casset, P. Ancey, and
A.M. Ionescu, in IEEE International Electron Devices
Meeting, 2005. IEDM Technical Digest, pp. 479–481 (2015).
11. F. Chen, H. Kam, D. Markovic, T.-J. K. Liu, V. Stojanovic,
and E. Alon, in 2008 IEEE/ACM International Conference
on Computer-Aided Design, pp. 750–757 (2008).
12. V. Pott, H. Kam, R. Nathanael, J. Jeon, E. Alon, and T.-J.
King Liu, in Proc. IEEE, pp. 2076–2094 (2010).
13. G. Pahwa, T. Dutta, A. Agarwal, S. Khandelwal, S.
Salahuddin, C. Hu, and Y.S. Chauhan, IEEE Trans. Electron Devices 63, 12 (2016).
14. S. Gupta, M. Steiner, A. Aziz, V. Narayanan, S. Datta, and
S.K. Gupta, IEEE Trans. Electron Devices 64, 8 (2017).
15. A.I. Khan, U. Radhakrishna, S. Salahuddin, and D. Antoniadis, IEEE Electron Device Lett. 38, 9 (2017).
16. A.I. Khan, K. Chatterjee, B. Wang, S. Drapcho, L. You, C.
Serrao, S.R. Bakaul, R. Ramesh, and S. Salahuddin, Nat.
Mater. 14, 2 (2015).
17. G. Catalan, D. Jiménez, and A. Gruverman, Nat. Mater. 14,
2 (2015).
18. A.I. Khan, D. Bhowmik, P. Yu, S.J. Kim, X. Pan, R. Ramesh,
and S. Salahuddin, Appl. Phys. Lett. 99, 11 (2011).
19. P.N. Kondekar and B. Awadhiya, in 2017 Joint IEEE
International Symposium on the Applications of Ferroelec-
22.
25.
26.
27.
28.
29.
30.
31.
32.
33.
34.
35.
36.
37.
38.
39.
40.
tric (ISAF)/International Workshop on Acoustic Transduction Materials and Devices (IWATMD)/Piezoresponse Force
Microscopy (PFM), Atlanta, GA, (2017), pp. 45–47.
H. Mehta and H. Kaur, Superlattices Microstruct. 111
(2017).
H. Mehta and H. Kaur, IEEE Trans. Electron Devices 65,
2699 (2018).
B. Awadhiya, P.N. Kondekar, and A.D. Meshram, Micro
Nano Lett. 13, 10 (2018).
B. Awadhiya, P.N. Kondekar, and A.D. Meshram, Superlattices Microstruct. 123, 306 (2018).
BSIM Group in UC Berkeley. http://wwwdevice.eecs.berkele
y.edu/bsim/?page=BSIM4.
D.J.R. Appleby, N.K. Ponon, S.K.K. Kwa, B. Zou, P.K. Petrov, T. Wang, N.M. Alford, and A. O’Neill, Nano Lett. 14, 7
(2014).
L.D. Landau and I.M. Khalatnikov, in Collected Papers,
L.D. Landau, Elsevier, pp. 626–629 (1965).
T.K. Song, J. Korean Phys. Soc. 46, 1 (2005).
C. Hu, S. Salahuddin, C.-I. Lin, and A. Khan, in 2015 73rd
Annual Device Research Conference (DRC), pp. 39–40
(2015).
Y. Li, K. Yao, and G.S. Samudra, IEEE Trans. Electron
Devices 64, 5 (2017).
A. Aziz, S. Ghosh, S. Datta, and S.K. Gupta, IEEE Electron
Device Lett. 37, 6 (2016).
S. George, S. Gupta, V. Narayanan, K. Ma, A. Aziz, X. Li, A.
Khan, S. Salahuddin, M.F. Chang, S. Datta and J. Sampson,
in Proceedings of the 53rd Annual Design Automation
Conference on—DAC’16, pp. 1–6 (2016).
S. Salahuddin and S. Datta, Nano Lett. 8, 2 (2008).
J. Müller, P. Polakowski, S. Mueller, and T. Mikolajick, ECS
J. Solid State Sci. Technol. 4, 5 (2015).
P. Polakowski and J. Muller, Appl. Phys. Lett. 106, 23
(2015).
K.D. Kim, M.H. Park, H.J. Kim, Y.J. Kim, T. Moon, Y.H.
Lee, S.D. Hyun, T. Gwon, and C.S. Hwaung, J. Mater.
Chem. C 4, 28 (2016).
A.I. Khan, U. Radhakrishna, K. Chatterjee, S. Salahuddin,
and D.A. Antoniadis, IEEE Trans. Electron Devices 63, 11
(2016).
L. Tu, X. Wang, J. Wang, X. Meng, and J. Chu, Adv. Electron. Mater. 4, 11 (2018).
A. Ortiz-Conde, F.J. Garcıa Sánchez, J.J. Liou, A. Cerdeira,
M. Estrada, and Y. Yue, Microelectron. Reliab. 42, 4
(2002).
T. Sakurai and A.R. Newton, IEEE J. Solid State Circ. 25, 2
(1990).
H. Mehta and H. Kaur, IEEE Trans. Electron Devices 66, 7
(2019).
Publisher’s Note Springer Nature remains neutral with
regard to jurisdictional claims in published maps and institutional affiliations.
Download