Download CNTMOSFETsrev5 - University of Maryland

Survey
yes no Was this document useful for you?
   Thank you for your participation!

* Your assessment is very important for improving the workof artificial intelligence, which forms the content of this project

Document related concepts

Quantum potential wikipedia , lookup

Renormalization wikipedia , lookup

Old quantum theory wikipedia , lookup

Introduction to gauge theory wikipedia , lookup

Aharonov–Bohm effect wikipedia , lookup

Quantum vacuum thruster wikipedia , lookup

Condensed matter physics wikipedia , lookup

T-symmetry wikipedia , lookup

History of quantum field theory wikipedia , lookup

Hydrogen atom wikipedia , lookup

Quantum electrodynamics wikipedia , lookup

Electron mobility wikipedia , lookup

Transcript
Quantum Modeling and Design of
Carbon Nanotube (CNT) Embedded
Nanoscale MOSFETs
Akin Akturk, Gary Pennington and Neil Goldsman
Department of Electrical and Computer Engineering
University of Maryland, College Park, MD 20742, USA
[email protected], [email protected], [email protected]
Abstract
We propose a novel MOSFET design that embodies single wall zig-zag semiconducting
Carbon Nanotubes (CNTs) in the channel. Investigations show that CNTs have high lowfield mobilities, which can be as great as 4x104 cm2/Vs. Thus we expect that MOSFET
performance can be improved by embedding CNTs in the channel. To investigate the
performance of a newly proposed CNT-MOSFET device, we develop a methodology that
connects CNT modeling to MOSFET simulations. Our calculations indicate that by
forming high mobility regions in the channel, MOSFET performance can be boosted.
However, barriers formed between the CNT and the silicon due to the variations of the
bandgaps and electron affinities can degrade MOSFET performance improvements. Our
calculations were obtained by building on our existing CNT Monte Carlo (MC) simulator
[1,2] and quantum based device solver [3,4].
Keywords: Single-Wall Zig-Zag CNT, CNT Mobility Model, CNT Embedded MOSFET
Design, Simulation.
1
I-Introduction:
As we approach the end of the semiconductor roadmap, investigators are exploring new
paradigms for electronic devices. Carbon nanotubes (CNTs) are being explored as a
structure that may play a leading role in future electronic systems [5]-[8]. CNTs are
planar graphite sheets (graphene) that are seamlessly wrapped into tubes. CNTs possess
favorable electrical characteristics and can be fabricated in dimensions as small as 8Å in
diameter. The electrical characteristics of CNTs vary with diameter and the wrapping
angle of the graphene [9]. Both the diameter and the wrapping angle can be described by
the tube’s fundamental indices (l,m) (Standard notation uses (n,m). However, l is used
here instead of n to avoid confusion with electron concentration.). Theory indicates that
CNTs can be metallic or semiconducting according to the fundamental tube indices (l,m),
with bandgap of the semiconducting tube depending on the CNT diameter. Analysis
shows semiconducting CNTs have very high low-field mobilities, with peak drift electron
velocities that can be as much as five times higher than that of silicon [1,2,10]. It has also
been shown that tubes can be doped by donors and acceptors [11]-[13]. Experiments and
calculations also indicate that CNTs may facilitate devices with large transconductances
and high drive currents [3,4], [10]-[21]. Experiments have demonstrated the viability of
CNT-based FETs [16,17], and CNT-SOI type MOSFETs [18,19]. Preliminary research
has been done to model and design CNT embedded bulk MOSFETs [3,4].
We investigate several CNT-MOSFET devices, similar to the one shown in Fig.
1. Our calculations indicate that CNT-MOSFETs can have improved device performance
over conventional MOSFETs [3,4]. To investigate the potential attributes of the new
design, we developed a methodology for modeling nanoscale CNT-MOSFETs. It
includes determination of the electrical characteristics of single wall zig-zag CNTs, and
the merging of the CNT results into our quantum device solver. To electrically
characterize the CNT, we developed a Monte Carlo (MC) simulator for CNTs. We first
calculate electron and phonon dispersion relations for single wall zig-zag CNTs with
different tube indices l. We then derive the selection rules and scattering matrix elements,
using Fermi’s Golden Rule. Using the MC results, we derive analytical models for CNT
parameters such as the mobility and the density of states. Once we obtain CNT
2
parameters, we import them to our quantum device solver. Our device solver is based on
the semiconductor equations, modified to account for CNT-silicon (CNT-Si) barrier
[20,21] and quantum effects. We solve these coupled equations on a mesh of our CNTMOSFET device. The solution gives results which include: CNT-MOSFET current
voltage curves, electron concentration profile in the bulk MOSFET, as well as in the CNT
enhanced channel.
II-CNT Simulator:
As mentioned above, we first employ a Monte Carlo (MC) simulator [1,2] to characterize
fundamental transport properties of CNTs, and then we incorporate these properties into
our device simulator. These properties include electron drift velocity versus electric field,
as well as CNT mobility for zig-zag single wall CNTs. (The zig-zag CNT is probably the
most studied semiconducting nanotube topology. Zig-zag CNTs have fundamental
indices (l,0), where l takes on integer values.) To obtain these properties, we begin with
the physical CNT system, where electrons are confined around the circumference, and
move relatively freely along the tube in the direction of the longitudinal axis. Therefore,
one can write the appropriate plane wave solutions that satisfy periodic boundary
conditions, distinguished by the quantum number β, for the given CNT circumference.
However along the tube, electrons are not confined. Thus the wavevector can be written
as follows:
k  kx x 
2
,
d
(   l ,
, l )
(1)
Here x is parallel to the tube axis, and  is the unit vector along the circumference.
Discrete values around the circumference, β, are bounded by the fundamental tube index
 l, to take advantage of the symmetry lines in the CNT Brillouin zone.
We consider one-dimensional semiclassical charge transport along the tube. If an external
field is applied in the longitudinal direction, electrons are accelerated along the tube until
they scatter with phonons. Here we consider scattering by optical, and acoustic inter and
intra subband, and intervalley phonons. Both scattering mechanisms are treated within
3
the deformation potential method using Fermi’s Golden Rule. For more details of our MC
simulations we refer the reader to [1,2].
A. CNT Bandstructure:
To account for the CNT related quantum effects in the CNT-MOSFET channel, we need
to determine the band-structure of the CNTs. Due to confinement, the bandstructure splits
into a system of subbands when graphene is wrapped into a CNT. Each of the subbands
has a characteristic effective mass, mobility and band energy minima. We determine the
energy levels of CNTs by applying zone-folding methods to graphene [9]. From the twodimensional graphene band diagram, we cut one-dimesional slices, whose numbers and
locations are set by the fundamental tube indices (l,m).
We use zone-folding of the graphene conduction band to calculate the electronic
bandstructure of the CNT. The following formula gives the energy dispersion for a zigzag CNT [9]:
 3

  
2   
E  3 1  4 cos 
ak x  cos 
  4 cos 

 l 
 l 
 2

(eV)
(2)
Where a (=2.46Å) is the lattice constant of two dimensional graphite. It has been shown
that this analytical formula agrees with experimental results.
To extract the pertinent information that can be easily integrated into our device
simulator, we transform Eqn. (2) into the following quadratic form which is consistent
with traditional semiconductor analysis:
2

k x2
  E  El  1   l  E  El 
2ml*

(3)
Here El is the conduction band minimum,  l is the non-parabolicity factor, and ml* is
the effective mass in subband β for a tube with index l. Conduction band minimum,
effective mass and non-parabolicity factor can be calculated using Eqn. (2) for different
4
subbands. For example, conduction band minimum for each subband can be found by
setting kx to zero in Eqn. (2):

   
El   3 1  2cos 

 l 

(eV)
(4)
Likewise, from the curvature of the energy dispersion relation given in Eqn. (2) around
kx=0 we obtain the following formula for the effective masses of each subband:
 El



l*
1eV
m

 0.0965 
mo
  
cos 

 l 
(5)
Our MC calculations indicate that only the lowest three subbands are effectively
populated for the electron transport calculations along the tube. Thus we only deal with
those subbands, which correspond to  βo,  2(l-βo), and  2(2βo-l), where βo is the
quantum number of the lowest energy subband. This subband is the one that is closest to
the K-point of the graphene Brillouin zone. From geometrical considerations, it is easy to
show that the value of βo for this subband is equal to 2l
3
rounded to the nearest integer.
Furthermore, the subbands are written in an order starting from the one with the lowest
energy minimum, followed by those with increasing minimum energy values. In Table 1,
we give a list of conduction band minima and effective masses for the first three
subbands of different tubes. In Fig. 2, we compare the l=10 CNT bandstructure found by
tight-binding method with the one calculated by Eqn. (3).
B. CNT Mobility Model:
We derived a mobility model based on our MC simulation results of drift velocity versus
electric field curves. These velocity versus field curves are plotted in Fig. 3 for CNTs
with tube indices ranging from 10 to 34, which correspond to diameters of 8Å to 27Å.
Simulation indicates that electron drift velocity first increases linearly with the applied
field, reaches a maximum, and then rolls off, showing a negative differential mobility
(NDM). We find that peak electron velocities are as much as five times higher than what
they are in silicon. Electrons reach velocities as high as 4.5x107cm/s in large diameter
5
CNTs (l=34), while the maximum velocity, which is approximately 3x107cm/s for l=10,
drops for smaller diameter tubes. However, these peak velocities in narrow tubes are still
larger than the corresponding values in other semiconductors. Calculated results also
show that the critical field, where we have the peak drift velocity, increases from 1kV/cm
to 10kV/cm as we reduce the tube diameter from 27Å to 8Å.
Figure 3 shows three main characteristics of the CNTs. First, CNTs attain drift velocities
larger than other semiconductors. Investigations show that this is due to decreased
scattering rates, which is a result of the quasi-one-dimensional system. Second, electrons
in large diameter tubes have higher velocities than the ones in small diameter tubes for a
given applied field, unless the applied field is too large. This leads to higher low-field
mobilities for larger diameter tubes. Analysis shows that this is due to lower effective
masses in the larger diameter tubes. Third, all CNTs show NDM. They are similar to
GaAs in that respect, where conduction band velocity of the first subband is larger than
that of the second.
We develop an analytical mobility model considering the two lowest subbands, which
dominate the conduction. By that means, we embed the effects of NDM in our mobility
model. We then express the final mobility using Mathiessen’s rule, as follows:
1
1
1


 (l , F ) 1 (l , F ) 2 (l , F )
(6)
Where 1 (l , F ) and 2 (l , F ) refer to the mobilities in the first and second subbands,
respectively. The mobilities are functions of fundamental tube index, l, and the electric
field, F. The mobility of the first subband is:
1 (l , F ) 
o (l )
F
1
Fc (l )
(7)
Here o (l ) is the low-field mobility, and Fc(l) is the critical electric field. The critical
electric field corresponds to the peak electron drift velocity. We have empirically
determined the following expressions for the low-field mobility and the critical field in
terms of tube index l:
6


o (l )  40l 2 1 
Fc (l ) 
 
l
2/3


1  64  
6
1  2  x10
l 
l 
3/ 2
(cm2/Vs)
(8)
(V/cm)
(9)
Here ρ=1-gcd(l+1,3) (= 0, -2), where gcd(l+1,3) is the greatest common divisor of l+1
and 3. The expression for low field mobility can be obtained from the familiar
expression o  q
m*
. Results from our previous work indicates that  is proportional
to l, and m* is inversely proportional to l [2,3], thereby giving the quadratic-type form of
Eqn. (8). We empirically write the mobility of the second subband as follows:
2 (l , F ) 
Vmax (l )

F 
F 1  

Fc (l ) 

(V/cm)
(10)
Here  is an empirical parameter which we find to have the value of 0.01. Vmax(l) is the
maximum drift velocity of the electrons, shown in Fig. 3. We find it to be the following
function of l:


Vmax (l )  1.5l1/ 3 1   x107
 2l 
(cm/s)
(11)
In Fig. 4, we show our calculated mobility versus field curves, using Eqns. (6)-(11). For
l=34, low-field mobility is as high as 4x104cm2/Vs, while for l=10 it is approximately
4x103cm2/Vs. Such high mobilities indicate that incorporating CNTs into MOSFETs may
yield high drive currents and transconductances.
We also found that high scattering rates help to validate the use a mobility model for the
CNTs we simulated that are about 0.14μm long and ranging in diameter from 8Å to 17Å.
Furthermore, mean free length versus electric field curves are concave down like drift
velocity versus electric field curves. In addition, the mean free paths (mfp) range from
approximately 10nm to a maximum of 100nm for the CNTs we use here. For smaller
diameters tubes (8Å) the mfp has a narrow peak value of approximately 30nm[1,2].
7
C. CNT Intrinsic Carrier Concentration:
To investigate the effects of embedding a CNT into a MOSFET, we developed a novel
device simulator. One of the fundamental quantities required by our CNT-MOSFET
solver is the CNT intrinsic carrier concentration. Thus we develop a methodology to
obtain the intrinsic concentrations of different tubes. We start from the parabolic
approximation to energy dispersion relation in Eqn. (3). The density of states for each
subband is zero for energies less than the energy minimum of that particular subband, and
becomes the following for energies greater than the subband energy minimum:
ml*
DOS 
l
2
2
(12)
E  E 
l

Using nondegenerate statistics and the zero energy point at the midgap, we get the
following expression for the intrinsic carrier concentration:

1
n 2
2
  DOS e
n 
kTml*
l
o
l
E
kT
dE
(13)
l
E
Or equivalently:
l
o

2 2
2
e

El 
1  t
kT
t
e dt
(14)
0
The integral in Eqn. (14) can be recognized as the gamma function with an argument
equal to ½, which makes the integral equal to  . The expressions in Eqns. (13,14) give
the one-dimensional carrier concentration. To obtain the intrinsic carrier concentration
per unit volume, we calculate the concentration that would arise by stacking quasi-2dimensional sheets of CNTs directly on top of each other to form a 3-dimensional volume
filled with nanotubes. Finally, we arrive at the following formula for the intrinsic carrier
concentration, which is a function of fundamental tube index l:
n 
l
o


1
2.49l


kTml*
2
2
8
2
e

El
kT
(15)
D. CNT Electron Affinity:
We last need the electron affinities of different size CNTs in addition to the intrinsic
carrier concentrations to include the effects of the CNT-Si barrier into the carrier
continuity equations. We use the bandgap of the CNT and the electron affinity of the
graphite to obtain the electron affinities of CNTs. We then calculate CNT affinities by
subtracting half the bandgap value of the lowest subband of the CNT from the electron
affinity of graphite, which is 4.4eV [22].
III-Quantum Device Simulator:
We develop a two-dimensional quantum device solver [3,4] based on the Poisson
equation and the modified semiconductor equations. We here take the invariance in the
width direction as retained by the introduction of tubes in the channel. Since CNTs in our
simulations have small diameters, the bending of the field around the tube is limited [23]
and the associated dielectric relaxation lengths are high enough to ensure smooth field
curves. The governing equations are listed below in the order of Poisson, quantum/CNTSi electron current continuity, and quantum/CNT-Si hole current continuity equations.
 2  
q

 p  n  D
(16)
n 1
 .J n  GRn
t q
(17)
p
1
  .J p  GR p
t
q
(18)
Here the newly introduced variables  , n-p, Jn-p, D and GRn-p are electrostatic potential,
electron-hole concentrations, electron-hole current densities, net dopant concentration,
and electron-hole Shockley-Hall-Read net generation-recombination rates, respectively.
We next define electron-hole current densities Jn-p as follows:
n
J n  qnn   QM  HS
  qn kT n
p
J p  qp p    QM  HS
  q p kT p
9
(19)
(20)
We here symbolize thermal voltage and electron-hole mobilities by TH and  n  p ,
respectively. We also introduce two additional effective potential terms QM and HS to
account for the quantum and the CNT-Si barrier effects, respectively. We next will
discuss how these two phenomena are taken care of by the effective potential terms
starting from the CNT-Si barrier effects.
Solution of the CNT-MOSFET system requires proper handling of two phenomena. The
first is the effect of the quantum well formed at the Si-SiO2 interface that causes band
splitting, thus lowering of the carrier concentration. Second one is the influence of the
barrier formed at the CNT-Si interface that results from the different bandstructures and
electron affinities of the CNT and silicon. A quantum well may also form at the CNT-Si
junction due to the bandoffsets.
A. Effective Potential due to CNT-Si Barrier
As an initial guess, we first solve our system without considering quantum confinement.
This translates to the coupled solution of Eqns. (16)-(20). At this stage, we also resolve
the effects of CNT-Si barrier, through the use of a revised current equation. To obtain this
form for the current, we start with the standard expression for the current as the gradient
of the quasi-Fermi potential:
J n  qnnn
(21)
Next, we introduce the familiar relationship between the quasi-Fermi potential, the
electrostatic potential and the intrinsic carrier concentration. However, for the CNT-Si
structure, the intrinsic carrier concentration, no, has spatial dependence.
n   
kT  n 
ln  
q  no 
(22)
We now multiply the numerator and denominator of the argument of the logarithm by a
constant which is equal to the intrinsic silicon carrier concentration to obtain the
following expression for n :
10

n    

kT  no
ln 
q  noSi
  kT  n 
ln  Si 
  
  q  no 
(23)
Substituting Eqn. (23) into Eqn. (21), we obtain the revised expression for the electron
current density:

kT  no
J n  qnn   
ln  Si
q
 no


   qn kT n

(24)
Here no is the intrinsic carrier concentration at a grid point on our device, and noSi is the
intrinsic carrier concentration of silicon. We note that no takes on either the intrinsic
carrier concentration of the CNT or the Si, depending on the location within the CNTMOSFET. The potential within the gradient of Eqn. (24) includes the electrostatic
potential and an effective potential due to the bandgap variations in CNT-Si structure
[24,25]. However, CNT and Si have also different electron affinities that would change
the potential barrier between these two materials. Thus we next introduce an additional
effective potential term that arises due to the effects of different electron affinities on
both sides of the CNT-Si barrier. In Eqn. (19), we sum the effects of the variations of the
bandstructure and the electron affinities in HS , which for electrons is defined as follows
[24]:
n
HS

1
kT  n 
   Si   ln  Sio 

q
q  no 
(25)
Where  is the electron affinity at a grid point on our device and is either equal to  Si
or  CNT . We subtract  Si from  because our reference material is Si as pointed out in
Eqn. (23). With the addition of the new term that incorporates different electron affinities,
electron current density becomes:

1
kT  no   
J n  qnn         Si  
ln 
    qn kT n

q  noSi   
q

(26)
We can also apply the same arguments to holes. Thus one would find the corresponding
effective potential expression for holes as written below:
p
HS

1
kT  n 
  EG   Si  EGSi   ln  Sio 

q
q  no 
11
(27)
Here bandgap EG, like  and no, refers to the same material in space. It takes on the
bandgap value of either the CNT or the Si depending on the location within CNTMOSFET.
B. Effective Potential due to Quantum Confinements:
Investigations show that carrier confinement at the Si-SiO2 interface and the CNT-Si
barrier can significantly reduce the carrier concentration adjacent to the interface [25][29]. In addition, the potential well formed at the band discontinuities between the CNT
and Si can result in confinement and band-to-band tunneling effects. To incorporate these
quantum effects in our device model, we use density gradient formalism. The density
gradient theory is based on an approximate many-body quantum theory [26]. It has been
shown that the density gradient theory resolves the effects of the MOSFET channel
confinement [27,28], band-to-band and source-to-drain tunneling [28,29]. In this
formalism, quantum effects are included by the introduction of an effective potential term
that is proportional to the gradient of the electron density. We express this potential,
which is QM in Eqn. (19), as follows [26]-[30]:
QM 
2 2
12q n
 1 2 n 1 2 n 



2
m y 2 
 m x
(28)
Here x is parallel to the MOSFET channel and tube axis, and y is normal to x. We next
use the effective mass of the Si or the CNT depending on the direction and location.
We treat the quantum induced effects in a manner that is analogous to the formation of
position dependent heterostructures in the quantum well. Thus we sum quantum effects in
the carrier concentration term, where quantum confinement is reflected as bandgap
broadening or lowering. This methodology introduces an extra potential term as follows:
kT  nQM
ln 
q  nCL
where
12

,

(29)
nQM
nCL
2

 1 2 n 1 2 n  
 exp 

.
 6kT n  m x 2
m y 2  


(30)
Here subscripts refer to quantum (QM) and classical (CL) solutions. We next incorporate
the term in Eqn. (30) into the current equation to account for quantization effects on
transport:

1
kT  no  nQM
J n  qnQM n        Si  
ln  Si 

q
q
 no  nCL

 
    qn kT nQM
 
(31)
Using a combination of numerical methods, we finally solve our coupled quantum
semiconductor Eqns (16)-(20) along with Eqns. (27,28), for the electrostatic potential,
quantum/CNT-Si electron concentration, and quantum/CNT-Si hole concentration for the
CNT-MOSFET. Once the aforementioned variables are determined, we then use them to
calculate the current-voltage characteristics of the CNT-MOSFET. We solve the system
numerically using the overall algorithm given in Fig. 5.
IV-Simulation Results:
We applied our modeling methodology to simulate a 0.15μm well-tempered CNTMOSFET [31]. We first simulated CNT-MOSFETs with a single layer of CNT in the
MOSFET channel parallel to the interface as illustrated in Fig. 1. Then parameter we
investigated in these simulations was the effect of different diameter tubes. We next
investigated how incorporating additional layers of 8Å tubes affects the device
characteristics.
In Fig. 6, we show our calculated electron concentration in the vertical direction of the
MOSFET channel, starting from the Si-SiO2 interface. We applied 1.5V to the gate
terminal, and grounded others. CNT-MOSFET contains one layer of tube. The device
with the medium diameter tubes (d=13 Å) shows high concentrations in the channel. The
abrupt change in the carrier concentration can be attributed the differences in the
conduction band offset between the CNT and Si. We associate this with the high intrinsic
carrier concentration and lower work function (compared to Si) of the larger diameter
tubes which attract electrons onto itself even in the absence of a gate field. On the other
13
hand, the intrinsic carrier concentration of the d=8Å CNT is close to that of Si, and has a
higher work function. Thus the potential well is formed on the tube which in turn pushes
electrons away from the channel of this CNT-MOSFET. Thus, the larger diameter CNTs
appear to be likely to sustain large transconductances. We next investigate whether the
band offsets between the wider tubes and silicon appear to negate the potential
improvement of higher electron concentration in the channel of the larger diameter tube
CNT-MOSFETs. Therefore we obtain the current-voltage characteristics of the 0.15μm
CNT-MOSFETs in the subthreshold, linear and saturation regions. In Fig. 7a, we
compare the drain current density versus applied drain voltage curves for four MOSFET
configurations. One set of curves is for the conventional MOSFET without any CNTs in
the channel. The other three sets of curves are for the single layer CNT-MOSFETs with
small (8Å), medium (13Å) and large (17Å) diameter CNTs in the channel, just below the
SiO2. We find that for high bias conditions, CNT-MOSFETs utilizing larger diameter
tubes attain higher drive currents than the ones having the small diameter tubes, followed
by the conventional MOSFET. One of the main differences in performance within CNTMOSFETs can be attributed to the height of the barrier formed at the CNT-Si junction.
The smaller diameter tubes have less barrier height offset since their intrinsic carrier
concentration is closer to that of silicon. However the small diameter tubes form a
potential well at the channel unlike the larger diameter tubes that attract more electrons as
the diameter gets bigger. The CNT-MOSFETs have improved drive current
characteristics over the conventional MOSFET. We attribute these higher currents to
larger channel electron concentrations shown in Fig. 6 and larger mobility values in
CNTs. However the large diameter tube CNT-MOSFET behaves more like a resistor with
a low output resistance due to the bandoffets and high mobility. In addition the small
diameter tube CNT-MOSFET has a jump in its current drive around VDS=0.6V where the
electron concentration on the tube suddenly jumps from the levels shown in Fig. 6 (1016)
to higher values (1018) indicating that new subbands are populated on the tube. We show
the subthreshold characteristics of the aforementioned CNT-MOSFETs in Fig. 7b. The
small diameter tube CNT-MOSFET has a steep subthreshold slope (like the conventional
device) with the least leakage level and higher drive currents compared to the
conventional device for high gate biases. We attribute this to the bandoffset and high
14
mobility associated with the small diameter CNTs. Additionally the small diameter CNTMOSFET shows negative differential transconductance. We associate this with the
occupation of new subbands on the tube with higher gate bias. For the same bias range,
larger diameter tube CNT-MOSFETs have a much higher leakage level which gets worse
as drain bias increases. However on/off current ratio is still on the order of a thousand,
which should enable their use as FETs but may limit their low power applications. We
attribute this to the bandoffsets and high mobility of the larger diameter tubes.
We next investigate ways to increase the electron concentration in the channel of the
small diameter tube CNT-MOSFET to achieve even higher current drives. The small
diameter tube device already has improved subthreshold characteristics, which are mainly
controlled by the bandoffsets at the drain and source sides. However drive current is
controlled by the gate via the electron channel formed in the CNT-MOSFET. Since
electron concentration is low on the tube due to confinement, we add extra layers of
CNTs in the vertical channel direction to increase the physical size of the well. (The
length of the tube is still in the direction of the channel.) Therefore, more electrons are
enabled to fit in the well. In Fig. 8, we show the electron concentration in the channel of
the small diameter tube CNT-MOSFET for various numbers of vertically stacked CNT
layers. We observe that the confinement effects are less pronounced as the number of
layers increases from one to three. This enables the peak electron concentration to be on
the CNTs, with a highest level reached for the three layered device. Therefore we
expected that this would have been mirrored in the drive current capabilities. We show
the current curves for high gate bias in Fig. 8a, where the highest current is supplied by
the three layered CNT-MOSFET. Additionally, the jump in the current drive of the one
layered device becomes less pronounced as the number of layers increases. We associate
this with less confinement in a well with bigger dimensions, where most of the states are
already occupied. In Fig. 8b, we show the subthreshold characteristics of these CNTMOSFETs. Our calculated currents show performance improvements as the number of
layers increases.
15
V-Conclusion:
We propose and investigate a novel device structure that combines MOSFET technology
with CNT nanostructures. We report that the CNT-MOSFET device appears to yield
better performance than the conventional MOSFET. To analyze the new design, we
develop a methodology for modeling CNT-MOSFETs. We first employ MC techniques
to electrically characterize single wall zig-zag CNTs. We then derive analytical models
for important CNT parameters, including mobility and intrinsic carrier concentration. We
next develop a methodology for incorporating these CNT characteristics into a quantum
device solver. We use the solver to calculate the current-voltage characteristics of CNTMOSFETs, as well as internal dynamic variables (quantum/CNT-Si electron
concentration, electrostatic potential, etc.) Our new CNT-MOSFET simulator predicts
that the drive current of CNT-MOSFETs is higher than that of conventional MOSFETs.
Likewise in the subthreshold region, the narrow diameter tube CNT-MOSFET shows
similar performance compared to the conventional device. Thus we conclude that CNTMOSFETs employing lower diameter carbon nanotubes appear to exhibit improved
capabilities and therefore may represent a new paradigm for devices in the 21st century.
16
VI. References
[1] G. Pennington, N. Goldsman, “Semiclassical Transport and Phonon Scattering on
Electrons in Semiconducting Carbon Nanotubes,” Phys. Rev. B, vol. 86, pp. 45426-37,
2003.
[2] G. Pennington, N. Goldsman, “Monte Carlo Study of Electron Transport in a Carbon
Nanotube,” IEICE Trans. Electron., vol. E86-C, pp. 372-8, 2003.
[3] A. Akturk, G. Pennington, and N. Goldsman, “Modeling the Enhancement of
Nanoscale MOSFETs by Embedding Carbon Nanotubes in the Channel,” Third IEEE
Conf. on Nanotech., pp. 24-7, 2003.
[4] Akturk, G. Pennington, and N. Goldsman, “Numerical Performance Analysis of
Carbon Nanotube (CNT) Embedded MOSFETs,” SISPAD 2004.
[5] M. Lundstrom, “A Top-Down Look at Bottom-Up Electronics,” Symp. on VLSI Cir.,
pp. 5-8, 2003.
[6] H. S. P. Wong, “Field Effect Transistors - From Silicon MOSFETs to Carbon
Nanotube FETs,” MIEL 2002, vol 1, pp. 103-107, 2002.
[7] J. Guo, S. Datta, and M. Lundstrom, “Assesment of Silicon MOS and Carbon
Nanotube FET Performance Limits Using a General Theory of Ballistic Transistors,”
IEDM 2002, pp. 29.3.1-4, 2002.
[8] Ph. Avouris, “Molecular Electronics with Carbon Nanotubes,” Acc. Chem. Res., vol.
35, pp. 1026-34, 2002.
[9] R. Saito, M. S. Dresselhaus, and G. Dresselhaus, Physical Properties of Carbon
Nanotubes, Imperial College Press, London, 1998.
17
[10] T. Durkop, S. A. Getty, E. Cobas, and M. S. Fuhrer, “Extraordinary Mobility in
Semiconducting Carbon Nanotubes,” Nano Letters, vol. 4, pp. 35-9, 2004.
[11] R. S. Lee, H. J. Kim, J. E. Fischer, A. Thess, and R. E. Smalley, “Conductivity
Enhancement in Single-Walled Carbon Nanotube Bundles Doped with K and Br,”
Nature, vol. 388, pp. 255-7, 1997.
[12] L. Grigorian, G. U. Sumanasekera, A. L. Loper, S. Fang, J. L. Allen, and P. C.
Eklund, “Transport Properties of Alkali-Metal-Doped Single-Wall Carbon Nanotubes,”
Phys. Rev. B, vol. 58, pp. R4195-8, 1998
[13] V. Derycke, R. Martel, J. Appenzeller, and Ph. Avouris, “Controlling Doping and
Carrier Injection in Carbon Nanotube Transistors,” App. Phys. Lett., vol. 80, no 15, pp.
2773-5, 2002.
[14] Z. Yao, C. L. Kane, and C. Dekker, “High-Field Electrical Transport in Single-Wall
Carbon Nanotubes,” Phys. Rev. Lett, vol. 84, pp. 2941-4, 2000.
[15] J. Hone, M. Whitney, C. Piskoti, and A. Zettl, “Thermal Conductivity of SingleWalled Carbon Nanotubes,” Phys. Rev. B, vol. 59, pp. R2514-6, 1999.
[16] A. Bachtold, P. Hadley, T. Nakanishi, and C. Dekker, “Logic Circuits with Carbon
Nanotube Transistors,” Science, vol. 294, pp. 1317-20, 2001.
[17] R. Martel, V. Derycke, J. Appenzeller, S. Wind, and Ph. Avouris, “Carbon Nanotube
Field-Effect Transistors and Logic Circuits,” IEEE Design Automation Conf., pp. 94-8,
2002.
[18] S. J. Tans, A. R. M. Verschueuren, and C. Dekker, “Room Temperature Transistor
Based on a Single Carbon Nanotube,” Nature, vol. 393, pp. 49-52, 1998.
18
[19] S. J. Wind, J. Appenzeller, R. Martel, V. Derycke, and Ph. Avouris, “Vertical
Scaling of Carbon Nanotube Field-Effect Transistors using Top Gate Electrodes,”
Applied Physics Letter, vol. 80, no. 20, pp. 3817-9, 2002.
[20] J. Appenzeller, J. Knoch, and Ph. Avouris, “Carbon Nanotube Field-Effect
Transistors – An Example of an Ultra-Thin Body Schottky Barrier Devices,” Device
Research Conf., pp. 167-70, 2003.
[21] J. Appenzeller, J. Knoch, R. Martel, V. Derycke, S. Wind, and Ph. Avouris, “ShortChannel like Effects in Schottky Barrier Carbon Nanotube Field-Effect Transistors,”
IEDM 2002, pp. 285-8, 2002.
[22] A. Charlier, R. Setton, and M-F. Charlier, “Energy Component in a Lattice of Ions
and Dipoles: Application to the K(THF)1,2C24 Compounds,” Phys. Rev. B, vol. 55, no. 23,
pp. 15537-43, 1997.
[23] X. Wang, H-S. P. Wong, P. Oldiges, and R. J. Miller, “Electrostatic Analysis of
Carbon Nanotube Arrays,” Sispad 2003, pp. 163-6, 2003.
[24] Y. Leblebici, S. Unlu, S-M. Kang, and B. M. Onat, “Transient Simulation of
Heterojunction Photodiodes-Part I: Computational Methods,” Jour. of Lightwave Tech.,
vol. 13, no. 3, pp. 396-405, 1995.
[25] A. S. Spinelli, A. Benvenuti, and A. Pacelli, “Self-consistent 2-D Model for
Quantum Effects in n-MOS Transistors,” IEEE Elect. Dev., vol. 45, pp. 1342-9, 1998.
[26] M. G. Ancona, G. J. Iafrate, “Quantum Correction to the Equation of State of an
Electron Gas in a Semiconductor,” Phys. Rev. B, vol. 39, no. 13, pp. 9536-40, 1989.
19
[27] M. G. Ancona, “Equations of State for Silicon Inversion Layers,” IEEE Elect. Dev.,
vol. 47, pp. 1449-56, 2000.
[28] M. G. Ancona, Z. Yu, R. W. Dutton, P. J. V. Voorde, M. Cao, and D. Vook,
“Density-Gradient Analysis of MOS Tunneling,” IEEE Elect. Dev., vol. 47, no. 12, pp.
2310-9, 2000.
[29] M. G. Ancona, “Macroscopic Description of Quantum-Mechanical Tunneling,”
Phys. Rev. B, vol.42, no. 2, pp. 1222-33, 1990.
[30] J. R. Watling, A. R. Brown, A. Asenov, A. Svizhenko, and M. P. Anantram,
“Simulation of Direct Source-to-Drain Tunneling Using the Density Gradient Formalism:
Non-Equilibrium Green’s Function Calibration,” SISPAD’02, pp. 267-70, 2002.
[31] http://www-mtl.mit.edu/Well/device50/doping/sh50.analytic
20
Table 1: Band structure parameters for zig-zag CNTs.
( ) β
0.086
0.527
7
2
0.358
1.146
6
3
0.221
1.854
8
1
0.058
0.333
11
2
0.177
0.704
10
3
0.170
1.243
12
1
0.043
0.244
15
2
0.118
0.507
14
3
0.137
0.929
16
1
0.029
0.159
23
2
0.070
0.326
22
3
0.099
0.616
24
Subband
10
1
16
22
34
ml*
El (eV)
l (m=0)
mo
21
List of Figures
1. Simulated CNT-MOSFET device………………………………………………..23
2. Bandstructure for a zig-zag CNT with l=10……………………………………..24
3. Electron drift velocity versus electric field curves for zig-zag CNTs with different
tube indices………………………………………………………………………25
4. Calculated mobility versus electric field curves for zig-zag CNTs with different
tube indices ……………………………………………………………………...26
5. Coupled algorithm flowchart…………………………………………………….27
6. Calculated electron concentration in the vertical direction of the Si-SiO2 interface
for CNT-MOSFETs with different diameter CNTs……………………….……..28
7. Calculated current-voltage curves for CNT-MOSFETs with different diameter
CNTs ……………………………………………………………...……………..29
8. Calculated electron concentration in the vertical direction of the Si-SiO2 interface
for CNT-MOSFETs with varying number of layers in the vertical channel
direction………………………………………………………………..….……..30
9. Calculated current-voltage curves for CNT-MOSFETs with varying number of
layers in the vertical channel direction…………………………………………..31
22
Figure 1: Simulated CNT-MOSFET device.
23
Figure 2: Bandstructure for a zig-zag CNT with l=10. Solid and dashed lines refer to
non-parabolic (Eqn. 3) and tight-binding (Eqn. 2) bandstructures, respectively.
24
Figure 3: Electron drift velocity versus electric field curves for zig-zag CNTs with
different tube indices.
25
Figure 4: Calculated mobility versus electric field curves for zig-zag CNTs with different
tube indices.
26
Figure 5: Coupled algorithm flowchart.
27
Figure 6: Calculated electron concentration profile in the middle of the CNT-MOSFET
channel, for different diameter CNTs and VG=1.5V (VD=VS=0V), starting from the SiSiO2 interface and going down about 9nm.
28
a)
b)
Figure 7: Current-voltage curves for CNT-MOSFETs with different diameter CNTs.
Calculated currents are for a) VGS=1.5V and b) VDS=1.0V (Inset shows the local
maximum point for the d=0.8nm tube CNT-MOSFET around VGS=1.4V.).
29
Figure 8: Electron concentration profile in the middle of the CNT-MOSFET channel, for
different number of CNT layers in the vertical channel direction and VG=1.5V
(VD=VS=0V), starting from the Si-SiO2 interface and going down about 6nm.
30
a)
b)
Figure 9: Current-voltage curves for CNT-MOSFETs with CNTs of 0.8nm in diameter
and varying number of tube layers (planar CNT sheets) in the vertical channel direction.
Calculated currents are for a) VGS=1.5V and b) VDS=1.0V (Inset shows the local
maximum point for the one layered CNT-MOSFET around VGS=1.4V. Two and three
layered CNT-MOSFETs show a weaker local maxima around VGS=0.5V.).
31