Two-dimensional numerical modeling of lightly doped nano-scale double-gate MOSFET

9
Two-dimensional numerical modeling of lightly doped nano-scale double-gate MOSFET Deepanjan Datta, A. Ananda Prasad Sarab, S. Dasgupta * Department of Electronics Engineering, Indian School of Mines, Jharkhand, Dhanbad 826004, India Received 5 May 2005; received in revised form 24 July 2005; accepted 27 July 2005 Available online 9 September 2005 Abstract A two-dimensional numerical solution of electrostatic potential and electric field profiles are presented for lightly doped nano-scale Double-Gate Metal-Oxide-Semiconductor-Field-Effect-Transistor (DG-MOSFET). We have developed quasi-static (QS) model for evaluating bulk and inversion charges based on symmetric linearization model. We have also shown the non-quasi-static (NQS) effect on the charge due to a time varying gate voltage. It is seen that various symmetries of DG-MOSFET characteristics with respect to source/drain interchange is maintained in quasi-static as well as non-quasi-static version of the symmetrically linearized model. The variation of the threshold voltage with the varying width of the device is evaluated and presented. The results have been compared and contrasted with reported analytical model for QS condition for the purpose of verification of the model. The variation of threshold voltage along the width of the device is also predicted. This numerical model can be extended to analyze the transport phenomenon in sub 30 nm channel length DG-MOSFETs. q 2005 Elsevier Ltd. All rights reserved. Keywords: Nano-scale modeling; Double gate MOSFET; Quasi-static charge; Non-quasi-static charge; QM modeling 1. Introduction As CMOS technology progresses, device dimensions have been scaled to nano-scale regime [1]. Therefore, in future transistors may operate near their ballistic limit rendering it is important to understand the ballistic device physics. An accurate modeling of nano-scale DG-MOSFET is essential using 2D simulator to accomplish the perform- ance analyses because of expected 2D Short Channel Effect (SCE). In low drain bias 1D electrostatics is sufficient to explain the characteristics of DG-MOSFET. But, at high drain bias, important 2D electrostatic effects like drain induced barrier lowering (DIBL) and threshold voltage roll- off can be explained using 2D modeling of nano-scale DG-MOSFET. As CMOS scaling is approaching the limit imposed by gate oxide tunneling, scaling down DG MOSFET is becoming a subject of intense VLSI research [2]. The advantages for DG MOSFET are ideal 60 mV/Dec sub-threshold slope [3], scaling by silicon film thickness without high doping, etc. Aggressive scaling of oxide layer thickness is required to provide substantial current drive to suppress SCEs such as Drain Induced Barrier Lowering (DIBL). To predict qualitatively the short channel threshold-voltage (V t ) roll- off and DIBL, full numerical potential model is described which includes both symmetric and asymmetric short channel DG-MOSFET models. These numerical results are essential for predicting the transport phenomena in short channel devices. Chen et al. [4] developed a unique non- equilibrium Green’s function approach (S.E.M.C.) for determining the characteristics of DG-MOSFET, which provides a physically rigorous approach to quantum approach and real scattering process. In another paper [5], Winstead and Ravaioli, modeled a 25 nm MOS capacitor which includes the non-equilibrium ballistic transport with quantum correction. Villanueva and Gamiz [6] have investigated the role of centroid inversion layer in DG-MOSFET. Here, the magnitude of the inversion layer centroid has been analyzed in volume-inversion transistors, where the minority carrier distribution has been found to be Microelectronics Journal 37 (2006) 537–545 www.elsevier.com/locate/mejo 0026-2692/$ - see front matter q 2005 Elsevier Ltd. All rights reserved. doi:10.1016/j.mejo.2005.07.009 * Corresponding author. Tel.: C91 326 220 6866; fax: C91 326 221 0028 E-mail address: [email protected] (S. Dasgupta).

Transcript of Two-dimensional numerical modeling of lightly doped nano-scale double-gate MOSFET

Two-dimensional numerical modeling of lightly doped nano-scale

double-gate MOSFET

Deepanjan Datta, A. Ananda Prasad Sarab, S. Dasgupta*

Department of Electronics Engineering, Indian School of Mines, Jharkhand, Dhanbad 826004, India

Received 5 May 2005; received in revised form 24 July 2005; accepted 27 July 2005

Available online 9 September 2005

Abstract

A two-dimensional numerical solution of electrostatic potential and electric field profiles are presented for lightly doped nano-scale

Double-Gate Metal-Oxide-Semiconductor-Field-Effect-Transistor (DG-MOSFET). We have developed quasi-static (QS) model for

evaluating bulk and inversion charges based on symmetric linearization model. We have also shown the non-quasi-static (NQS) effect on the

charge due to a time varying gate voltage. It is seen that various symmetries of DG-MOSFET characteristics with respect to source/drain

interchange is maintained in quasi-static as well as non-quasi-static version of the symmetrically linearized model. The variation of the

threshold voltage with the varying width of the device is evaluated and presented. The results have been compared and contrasted with

reported analytical model for QS condition for the purpose of verification of the model. The variation of threshold voltage along the width of

the device is also predicted. This numerical model can be extended to analyze the transport phenomenon in sub 30 nm channel length

DG-MOSFETs.

q 2005 Elsevier Ltd. All rights reserved.

Keywords: Nano-scale modeling; Double gate MOSFET; Quasi-static charge; Non-quasi-static charge; QM modeling

1. Introduction

As CMOS technology progresses, device dimensions

have been scaled to nano-scale regime [1]. Therefore, in

future transistors may operate near their ballistic limit

rendering it is important to understand the ballistic device

physics. An accurate modeling of nano-scale DG-MOSFET

is essential using 2D simulator to accomplish the perform-

ance analyses because of expected 2D Short Channel Effect

(SCE). In low drain bias 1D electrostatics is sufficient to

explain the characteristics of DG-MOSFET. But, at high

drain bias, important 2D electrostatic effects like drain

induced barrier lowering (DIBL) and threshold voltage roll-

off can be explained using 2D modeling of nano-scale

DG-MOSFET. As CMOS scaling is approaching the limit

imposed by gate oxide tunneling, scaling downDGMOSFET

is becoming a subject of intense VLSI research [2].

0026-2692/$ - see front matter q 2005 Elsevier Ltd. All rights reserved.

doi:10.1016/j.mejo.2005.07.009

* Corresponding author. Tel.: C91 326 220 6866; fax: C91 326 221

0028

E-mail address: [email protected] (S. Dasgupta).

The advantages for DG MOSFET are ideal 60 mV/Dec

sub-threshold slope [3], scaling by silicon film thickness

without high doping, etc.

Aggressive scaling of oxide layer thickness is required to

provide substantial current drive to suppress SCEs such as

Drain Induced Barrier Lowering (DIBL). To predict

qualitatively the short channel threshold-voltage (Vt) roll-

off and DIBL, full numerical potential model is described

which includes both symmetric and asymmetric short

channel DG-MOSFET models. These numerical results

are essential for predicting the transport phenomena in short

channel devices. Chen et al. [4] developed a unique non-

equilibrium Green’s function approach (S.E.M.C.) for

determining the characteristics of DG-MOSFET, which

provides a physically rigorous approach to quantum

approach and real scattering process. In another paper [5],

Winstead and Ravaioli, modeled a 25 nm MOS capacitor

which includes the non-equilibrium ballistic transport with

quantum correction. Villanueva and Gamiz [6] have

investigated the role of centroid inversion layer in

DG-MOSFET. Here, the magnitude of the inversion layer

centroid has been analyzed in volume-inversion transistors,

where the minority carrier distribution has been found to be

Microelectronics Journal 37 (2006) 537–545

www.elsevier.com/locate/mejo

Fig. 1. Schematic of the Double Gate MOSFET under study.

D. Datta et al. / Microelectronics Journal 37 (2006) 537–545538

closer to the interfaces in volume inversion transistors and

therefore, some of the advantages assumed in the device

have been grounded. Ren and Lundstrom [7] have derived

the numerical simulation technique from a Schrodinger–

Poisson tool applied to DG MOSFET where essential

physics governing the transport phenomena is presented to

give the insight concerning the near-ideal features of

DG-MOSFET. A compact, physics based short channel

models of sub threshold swing and threshold voltage model

[8] is presented for undoped symmetric DG-MOSFETs,

which includes quantum-mechanical and fringe-induced

barrier lowering effects. Baccarani et al. [9] investigated the

electrical properties of the DG MOSFET, which turn out to

be very promising for device miniaturization below 0.1 mm.

A compact model which accounts for charge quantization

within the channel, Fermi statistics, and non-static effects in

the transport model was worked out. Liu and Hosokawa [10]

investigated in a fabrication process of vertical gate DG

MOSFET using ion-bombardment-retarded etching. The

fabricated vertical DG-MOSFETs clearly exhibit the unique

advantage of DG MOSFETs, i.e. high improvement of

short-channel effect immunity by reducing the channel

thickness. An analytical expression explicitly relating the

potential and the electric field at the oxide-semiconductor

interface of a symmetrical double-gate oxide-intrinsic

semiconductor-oxide was derived by Wang et al. [11].

In this paper, the ‘turn-on’ behavior of the system was

studied and an ‘extrapolated’ threshold voltage was

uniquely defined.

In steady path of shrinking the device geometry, many

compact models for the device are proposed. However,

these models have severe drawbacks during their develop-

ment. Models like SPICE level 1 and level 2 are highly

empirical and have constrained discontinuities in their

current and capacitance characteristics. These models did

not incorporate the physics of short channel effect correctly.

Other models like MISNAN model, BNR and the Siemens

model developed by Miura-Mattausch [12] do not have

these continuities under certain limited conditions, but

display unrealistic conductance behavior at the linear/

saturation transition boundary on the output (Id–Vd)

characteristics [13]. The common problems with models

like BSIM3v3 and EKV models are that they commonly

took source a reference point due to which the expressions

containing potential, terminal charges, and transconduc-

tance lost the essential symmetry along the channel.

In this paper, we have developed a quasi-static and non-

quasi-static profile of the bulk and inversion charges of

symmetrical as well as asymmetrical DG MOSFET with

low doping levels. This paper solved the complete boundary

value to obtain numerical solution of 2D potential and

electric field profile for symmetric and asymmetric

DG-MOSFETs. For the first time, the charge profile is

developed taking into account the symmetricity with respect

to source/drain interchange in the case of DG-MOSFETs.

The bulk and inversion charge densities in both QS and

NQS profile have shown the expected symmetry at zero

drain bias. The potential and electric field profile for the

whole active area of the device has been evaluated and

presented. We have also investigated the transient phenom-

ena in terms of current flow, where gate voltage varies with

time in gaussian manner. The results obtained on the basis

of a model have been compared and contrasted with respect

to reported analytical results.

2. Numerical modeling

A 2D numerical modeling scheme has been applied to the

DG-MOSFET. Fig. 1 shows the schematic diagram of a

lightly-doped DG-MOSFET under study. The origin has

been chosen to be the centre of the Si substrate at the

source–substrate boundary as shown in Fig. 1.

For lightly doped DG MOSFETs, both the fixed and

mobile charge densities are almost negligible. In both the

insulator and silicon regions, Poisson’s equation becomes

d2J

dx2C

d2J

dy2Z

q

3Siexp

qJðx; yÞ

kT

� �(1)

If we assume the source and drain junctions are abrupt,

the boundary conditions can be written as

Jð0; yÞZVbi; KtSi2!y!

tSi2

(2)

JðL; yÞZVbi CVds; KtSi2!y!

tSi2

(3)

J x;KtSi2Ktox

� �ZVgsKDF1; 0!x!L (4)

J x;W

2C tox

� �ZVgsKDF2; 0!x!L (5)

D. Datta et al. / Microelectronics Journal 37 (2006) 537–545 539

Here, Vbi is the built-in-voltage, Vds is the applied drain

voltage, tox is the oxide layer thickness and xSi is the

permittivity of the Si layer. K is the Boltzmann coefficient

and T is the room temperature (TZ300 K). Here, L is

the channel length and tSi is the thickness of the device. The

same voltage Vgs is assumed to be applied to both the

gates. DF1 and DF2 are the work function differences

between the gate electrode 1,2 and the intrinsic silicon.

Hence, DF1Z0 for mid gap work function gate in case of

symmetric DG-MOSFET. The electrostatic potential is

defined as the intrinsic potential at point (x,y) with respect to

the Fermi potential of nC source (at the conduction band

edge).

Results show that in the insulator regions, between

source/drain and gates, the potential is given as linear

function of y, when tox/tSi!1 [14]. The potential in the

insulator gap regions are obtained by suitable interpolation

and can be expressed as:

Jð0; yÞZVbi

yC tSi2C tox

� �tox

KðVgsKDF1Þ yC tSi

2

� �tox

;

KtoxKtSi2!y!K

tSi2

(6a)

JðL; yÞZ ðVbiCVdsÞyC tSi

2C tox

� �tox

KðVgsKDF1Þ yC tSi

2

� �tox

;

KtoxKtiSi2!y!K

tSi2

(6b)

Jð0; yÞZKVbi

yKtSi2Ktox

� �tox

CðVgsKDF2Þ yKtSi

2

� �tox

;

tSi2!y!

tSi2C tox

(6c)

JðL;yÞZKðVbiCVdsÞyKtSii

2Ktox

� �tox

CðVgsKDF2Þ yKtSi

2

� �tox

;

tSi2!y!

tSi2Ctox

(6d)

In a lightly doped (or, practically undoped body)

DG-MOSFET we generally neglect hole concentration.

We have also taken aluminum (Al) as the gate metal, so

work functions of the both the gates are taken as (DF1ZDF2Z) 0.2 V for symmetric DG-MOSFETs.

A 2D numerical solution to the Poisson equation is

composed of Nx!NY potential values at each lattice node.

To attain Nx!NY unknowns, we have taken the finite

difference method and a uniform grid is superimposed on

the geometry of the channel with mesh spacing of hZ1.6!10K10 m and kZ1.6!10K10 mm along x and y-directions,

respectively. This means that at any point on the grid, using

classical discretization formula we can write

Ji;j Z1

2½JiK1;j CJiC1;j C ðh2=k2ÞðJi;jK1 CJi;jC1Þ�

!h2k2

h2 Ck2K

qh2

3Siexp

qJi;j

kT

� �(7)

where Ji,j and JiG1,jG1 are the surface potential points in

the channel. We can formulate two continuity equations at

two Si–SiO2 interfaces as:

dJ

dyyZtSi=2

ZKxSi

xOxtOxðJðx; tSi=2ÞKVgs CVfb1Þ

���� (8a)

dJ

dyyZKðtSi=2Þ

ZxSi

xOxtOxðJðx;KtSi=2ÞKVgs CVfb2Þ

���� (8b)

Here, permittivity of the oxide layer is taken as xOx. Vfb1

and Vfb2 are the flat-band voltages of the two gate sides. The

potential values in the Si layer is obtained for the solution of

Poisson’s equation under specified boundary conditions by

finite difference technique and subsequently utilized for

computing partial derivatives of J(x, y) with respect to y

and x to obtain transverse and longitudinal electric field,

respectively.

3. Quasi-static charge model for DG-MOSFET

The charge model allows one to explain the closed form

expressions for both the drain current and the terminal

charges as shown by Brews [15] and Ward et al. [16]. The

formulation by Ward et al. is however more involved than is

desirable in compact MOSFET models. Based on the

potential profile of symmetric DG-MOSFETS, we have

developed a linearized model for both the bulk charge and

the inversion charges, considering the requirements of

‘Gummel Symmetry’.

Defining normalized inversion charge density qi as, qiZQi/Cox; drain current can be expressed as

Ids ZKWCoxVqi CWCoxD dqi=dx (9)

where Qi is the inversion charge density and Cox is the

capacitance of the oxide layer. D is the diffusion co-efficient

in the steady-state drift-diffusion equation verifying the

continuity of the drain current. For the carrier velocity

expression, it is important to note that, it is symmetric with

respect to zero fields point and are continuous for all fields

within the depletion region of the device. We can take

carrier velocity as [14]:

vZmðEÞdJ

dx

� �= 1C

mðEÞ

vsat

dJ

dx

� �2� 1=2

(10)

vSAT is taken as 105 m/s [14] and m(E) is the transverse

direction dependent electric field which can be written as

[17]

D. Datta et al. / Microelectronics Journal 37 (2006) 537–545540

1

mðEÞZ

1

mc

C1

mph

C1

msr

(11)

where mc, mph and msr are coulomb scattering limited

mobility, phonon scattering limited mobility and surface

roughness scattering limited mobility, respectively.

(i) Coulomb-scattering limited mobility. Due to the

presence of fixed oxide charge, interface trapped charges

and the ionised impurities the carriers in the channel get

scattered due to coulombic perturbation, as a result the

mobility of the carriers in the surface channel decreases.

The coulomb scattering limited mobility can be modelled

as [17]

mc Z aK11 T (12)

where a1 is an extractable parameter and T is the absolute

temperature.

(ii) Phonon-scattering limited mobility. The phonon

scattering limited mobility can be modelled as [17]

mph Z aK12 TKnE

K1=geff (13)

where g and n are empirical constant. a2 is another

extractable parameter [17]. Here, Eeff is

Eeff Zq 1

2Ns CNdepl

� �3Si

(14)

Ns is the inversion layer carrier density, Ndepl is the

depletion region carrier density.

(iii) Surface roughness scattering limited mobility. This

is due to the deviation of the surface from an ideal plane.

The surface roughness scattering is independent of

temperature and is dominant at high electric fields. The

surface roughness scattering limited-mobility can be written

as [17]

msr Z aK13 EK2

eff (15)

where a3 is an extractable parameter.

From the MOS physics, normalized inversion charge

density can be expressed as a function of potential within

the channel and is written as

qiðJÞZVGSKVfbKJðxÞKQb=Cox

where Qb represents the bulk charge density. In order to

implement the approach to linearization, one may represent

the bulk charge as

QbðJÞZQbS CQbD

2

� �Cga JK

JSS CJSD

2

� �(16)

where Qbs and QbD are the bulk charge at source and drain,

respectively. Source terminal potential is denoted by Jss

and drain terminal potential by JSD; g is the Gummel

linearization factor and a is given by

aZ1

2

dq

dJ

� �JZJSS

Cdq

dJ

� �JZJSD

� (17)

Flexibility of the symmetric linearization is illustrated in

[18]. The linearization preserves essential symmetry at the

zero drain bias and it verifies almost symmetric character-

istics of the charge profile also at zero drain bias,which is also

true for inversion charge density. In themodels likeBSIM3v3

(Berkeley) or, EKV (Enz. et al.) [19], the bulk charge

or, threshold voltage is expressed in one-sided manner,

i.e. they took source as the reference point. This is the reason

for the discontinuity of the terminal charges, transconduc-

tance and transcapacitance profile in above models.

Likewise, we may represent the inversion charge, which

is a function of potential as perfect symmetrical and

linearized in nature and can be given as

qiðJÞZqiSCqiD

2

� �Kð1CgaÞ JK

JSSCJSD

2

� �(18)

where qis and qiD are the exact values of the normalized

inversion charge density at the two terminals along the

channel. Thus the channel current IDS can be written as [13]

IDS

ðL0

1CmðEÞ

vSAT

dJ

dx

� �2� 1=2

=meffdx

Z tSiCox K

ðJSD

JSS

qi dJCVt

ðqiDqiS

dqi

8><>:

9>=>; (19)

or

IDSZmðEÞCoxðtSi=LÞfðqiSCqiDÞðFDKFSÞ=2CVtðqiSKqiDÞg

1C meff

vSAT

ðFDKFSÞllin

� �2� 1=2

(20)

where Fs and FD be the value of the potential at the source

and drain ends and Vt is the surface potential at the onset of

strong inversion.

Defininggeff as effective value of body effect coefficientwe

can obtain bulk charge and normalized inversion charge as

QbS ZgeffCox

ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiJSSCVt e

KJSS=VtKVt

q(21)

QbS Z qiS ZVgsKVfbKJSSKQbS=Cox (22)

Similarly, we can get the expressions for bulk charge and

normalized inversion charge for drain end, which are used to

evaluate the profile of bulk charge and inversion charge along

the channel length. At this point brief discussion of the physics

behind the drain saturation voltage is necessary. With a fixed

gate voltage Vg, if the drain voltage is increased, the inversion

charge density at the drain decreases. As qiD decreases, the

potential along the channel adjusts to increase the carrier

velocity at the drain so that the current flow is maintained

along the channel. However, there is a limit of the decrement

of qiD because maximum velocity the carrier can achieve is

vSAT. Once applied drain voltage is large enough to reach

D. Datta et al. / Microelectronics Journal 37 (2006) 537–545 541

the maximum carrier velocity, qiD is pinned to a minimum

value. Correspondingly, potential at drain has reached to

maximum.

With the help of Eqs. (11) and (13), we can determine the

bulk and normalized inversion charge. The symmetricity is

achieved by taking the average of the two terminal values.

Thus, in quasi-static charge model, the advantage of

symmetric linearization method is that explicit form of

J(x) dependence is available.

Fig. 2. Variation of drain current with gate voltage in the sub-threshold

region for symmetric DG-MOSFETs for the proposed numerical model and

reported 2D analytical model [14].

4. Non-quasi-static charge-sheet model

In QS approach described above, the channel charge is

assumed to achieve equilibrium, once biases are applied,

thus the finite charging time of the carriers in inversion layer

is ignored. This gives erroneous results for signals, with rise

or, fall time comparable to or smaller than the channel

transit time. Moreover, none of the QS approaches predict

correctly, the terminal currents taking into account the

inertia in the charge transport of carriers from source to

drain and the charge distribution transient when rapidly

varying voltages are applied to the device terminals. In NQS

model the channel charge is not only the function of the

terminal voltages, but also is a function of time. It can be

expressed in a generalized form

QðtÞZQðt;VSðtÞ;VDðtÞ;VGðtÞ;VBðtÞ: (23)

where Vs(t) is the source voltage varying with time VD(t) is

the time-varying drain voltage and VG(t) is the gate voltage

varying with time.

Differentiating Eq. (23) with respect to time, we get

dQ

dtZ

vQi

vtC

XjZS;D;G;B

vQi

vVj

dVj

dt(24)

The charge sheet NQS is formulated in [20] and it

requires the solution of the following differential equation

vqivt

CmðEÞv

vx

qidqidJ

KVt

!vqivx

" #Z 0 (25)

obtained by combining the drift-diffusion expression with

current continuity equation.

The MOSFET inversion layer charge in NQS model can

be expressed in simple form in terms of the surface potential

using the charge-sheet approximation. The normalized

inversion charge per unit area can be expressed as a

function of surface potential along the channel length as

qiSðtÞZVgsðtÞKVfbKJSSKQbSðtÞ=Cox (26)

qiDðtÞZVgsðtÞKVfbKJSDKQbDðtÞ=Cox (27)

where

QbSðtÞZgeffCox

ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiJSS CVt e

KðJSS=VtÞKVt

q(28)

QbDðtÞZgeffCox

ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiJSD CVt e

KðJSD=VtÞKVt

q(29)

Here, we have taken the time-varying gate voltage to

achieve the NQS model, where, gate voltage is varying in

gaussian manner with time. We can achieve the required

symmetric linearization by using Eqs. (21) and (22), where

inversion and bulk charges are varying with time in the

manner given in Eqs. (26)–(29) and potential of the

symmetric DG MOSFET can also be varied with time

determined by the boundary conditions, where, gate voltage

is increasing in gaussian manner with time for getting drain

current profile in NQS model. The applied gate voltage is

varied as 1D gaussian profile and is given by, VgsZ0.02!exp(t2). With the variation of gate voltage we have

structured the profile for bulk and normalized inversion

charges which are also satisfying the symmetric lineariza-

tion technique to get symmetricity along the channel length.

The verification of the ‘Gummel Symmetry’ is one of the

important criterions to develop continuous and error-free

profile of the drain current also in NQS method.

5. Results and discussion

5.1. DG MOSFETs

Simulations have been carried out for nano-scale DG

MOSFET with channel length 10 nm and channel width

10 nm. The oxide layer thickness has been taken to be

0.5 nm. For the purpose of verification, the results obtained

as the basis of our model is compared and contrasted with

reported results [14] for a symmetric DG-MOSFET and is

shown in Fig. 2. The figure shows the variation of drain

current with applied gate voltage for our model in QS

domain and also the result obtained by Taur et al. [14] for

channel length of 24 and 200 nm having oxide layer

thickness of 1.5 nm in sub-threshold region. The applied

drain voltage has been taken to be 1 V. The agreement is

fairly accurate in the sub-threshold region of the device

which proves the validity and accuracy of the proposed

numerical model for DG-MOSFET. The slight deviation in

Fig. 3. Electrostatic potential contour plot for symmetric DG-MOSFETs for

VdsZ0 V, VgsZ0.02 V, tSiZ10 nm, LZ10 nm and toxZ0.5 nm.

Fig. 5. Variation of the threshold voltage with channel width for two

applied drain voltages, viz. 1 and 2 V.

D. Datta et al. / Microelectronics Journal 37 (2006) 537–545542

the extremely short channel device is due to omission of the

higher order terms in the analytical model proposed in [14].

Electrostatic potential contours are calculated at a drain

voltage of 0 V and shown in Fig. 3 for a channel length and

width of 10 nm, respectively. At the Si–SiO2 interface, the

surface potential is approximately equal to 0.13 V near

the source/drain end. The potential profile (in 3D) in the

insulator region is given by the interpolation formula as

described earlier, i.e. we get a linear profile inside the oxide

layer. In the depletion region the potential profile along the x

direction first decreases, reaches a minimum almost in the

middle of the channel and then again increases and reaches a

value at the drain end which is governed by the boundary

conditions. At zero drain bias, we get a symmetric 3D

profile as shown.

The electric field profile is shown in Fig. 4. It can be seen

from the figure that at the middle of the channel length as

one moves from one gate to another, the electric field first

decreases, reaches to a value equal to zero and then goes to a

negative value. It can be further observed that at the center

of the channel width as one move from source to drain

electric field is almost zero. This may be inferred from the

fact that at the center of the width, both the gate has

Fig. 4. Electric field profile for symmetric DG-MOSFETs with VdsZ0 V,

VgsZ0.02 V.

opposing effects on the potential profile for symmetrical

DG-MOSFET.

Fig. 5 shows the variation of the threshold voltage of the

DG-MOSFET with varying channel width for two applied

drain-to-source voltage. It can be seen from the figure that as

the channel width increases, the threshold voltage initially

increases up to a width of 3 nm and then remains constant up

to about 7 nm, subsequently decreases. It can be further

observed that at a particular width of the device, the

threshold voltage is lowered for a higher applied drain to

source voltage. Here, we have calculated the profile of

inversion charge density (normalized with respect to Cox)

shown in Fig. 6 and bulk charge profile shown in Fig. 7 at

different drain voltages. The profiles show symmetricity at

zero drain bias with respect to source/drain interchange

ensuring ‘Gummel Symmetry’. At any cross-section of the

device the sum of the inversion charge and the bulk charge

density will be zero, which reveals that the described

charge-sheet model is continuous, error-free and valid for

any short channel device. As seen in Fig. 7 the bulk charge

for three different drain voltages shows same value at a

distance approximately 8 nm away from the source end.

This is due to DIBL effects at the drain end. Fig. 8 shows the

variation of the drain current with gate voltage for three

devices of channel lengths 5, 10 and 15 nm, respectively.

The current level corresponds to a gate voltage 0.2 V below

threshold making it well into sub-threshold region.

Fig. 6. Variation for normalized inversion charge with distance along the

channel for three different values of drain voltages, viz. 0, 1 and 2 V for

quasi-static case.

Fig. 7. Variation of bulk charge with distance along the channel for the

three different values of drain voltages, viz. 0, 1 and 2 V for quasi-static

case.

Fig. 8. Variation of drain current density with gate voltage for three

different channel lengths, viz. 5, 10 and 15 V at VdsZ1 V in the threshold

region.

Fig. 10. Variation of normalized inversion charge in the symmetric DG-

MOSFET in non-quasi-static case.

D. Datta et al. / Microelectronics Journal 37 (2006) 537–545 543

The applied drain voltage is taken as 1 V. It can be seen that

the drain current is maximum for LZ5 nm for a particular

gate voltage. It can be further inferred that the transconduc-

tance is maximum at LZ5 nm.

In NQS profile, the gate voltage is varied as 1D gaussian

profile to get the variation of charges along the device

dimension. We have calculated the bulk charge as shown in

Fig. 9. It can be observed from figure that as one move from

source to drain, the bulk charge decreases to a minimum

value and subsequently increases. The symmetricity of

Fig. 9. Three-dimensional bulk charge profile in non-quasi-static case with

application of gate voltage varying with time in gaussian manner.

the bulk charge across the distance of the channel also

proves the correctness of the profile. The normalized

inversion charge profile as shown in Fig. 10 shows the

first sudden decrease of inversion charge and then it will

increase almost in a linear fashion and then acquires a

almost constant value. The decrease in normalized charge is

counterbalanced by increase in bulk charge so that at any

cross-section through the device at any instant of time the

sum of the inversion charge and the bulk charge will be

zero. It indicates the validity of the profile as in the case of

quasi-static charge distribution.

5.2. Asymmetric DG MOSFETs

For asymmetric MOSFET two work functions at the gate

ends are not equal. Therefore, one channel turns on first.

Similar to symmetric DGMOSFET, the 3D potential profile

can be obtained from the numerical solution for asymmetric

DG MOSFETs and is shown in Fig. 11.

Fig. 11. 3D Electrostatic potential plot for asymmetric DG-MOSFETs.

Here, work-function of the two gates are DF1Z0.2 V and DF2ZK0.2 V,

VdsZ0 V, VgsZ0.02 V.

Fig. 12. 3D electric field profile for an asymmetric DG-MOSFET for

applied drain voltage of 0 V.

Fig. 14. Variation of normalized inversion charge for four different values

of applied drain voltages, viz. 0, 0.2, 0.4 and 0.6 V for asymmetric DG-

MOSFETs for quasi-static case.

Fig. 15. Variation of drain current with gate voltage for an asymmetric

DG-MOSFET three applied drain voltages, viz. 1.5, 1 and 0 V.

D. Datta et al. / Microelectronics Journal 37 (2006) 537–545544

Here, we have taken DF1ZKDF2Z0:2 V, i.e. DF1,

2s0. For asymmetric DG MOSFETs, the minimum

potential is located not exactly at the middle of the channel

length but it is closer to one gate. Therefore, the minimum

potential in the asymmetric DG MOSFET is controlled

mainly by the gate than by the source/drain. The minimum

potentialJmin is located at the (xczL/2, ycztSi/2), where xcis the position, where J has the minimum value.

Fig. 12 shows a 3D electric field profile for an

asymmetric DG MOSFET for an applied drain-to-source

voltage of 0 V. It can be seen from the figure that at any

point in the channel, as we move from top gate to the bottom

gate of the DG-MOSFET, there is a decrease in the absolute

value of the electric field initially decreases and reaches a

value of almost zero at the middle of the channel width and

subsequently increases as one moves towards the bottom

gate of the DG-MOSFET. It can be further observed that the

electric field reversal takes places at the middle of the

channel width. This is due to the fact that, the potential

shows a reversal at the middle of the channel width due to

the configuration of the symmetric DG-MOSFET.

The bulk charge profile is shown in Fig. 13 and

normalized inversion charge shown in Fig. 14 shows the

Fig. 13. Variation of bulk charge with distance along the channel for four

different values of applied drain voltages, viz. 0, 0.2, 0.4 and 0.6 V for

asymmetric DG-MOSFETs in quasi-static case.

general trend of the charge models for four different drain

voltages. In the bulk charge profile the charge first decreases

along the distance along the channel, reaches a minimum

and then increases to a maximum. The normalized inversion

charge shows the reverse behavior of the bulk charge profile

to maintain the total charge zero at any cross-section of the

device. Fig. 15 shows the variation of drain current with

applied gate voltage for three different drain-to-source

voltage for an asymmetric DG-MOSFET. It can be seen

from the figure that the drain current shows almost a linear

increase with an increase in the applied drain voltage.

Further, at a particular gate voltage, the drain current is

larger for a larger drain voltage. It can be further seen that

the transconductance of the device for an applied drain

voltage of 1.5 V is maximum amongst all the three applied

voltages.

6. Conclusion

A full 2D numerical modeling of symmetric and

asymmetric DG MOSFET is described taking into con-

sideration quasi-static and non-quasi-static charge model.

This model highlights the nature of electrostatic potential

and transverse electric field, which is necessary to calculate

the drain current as well as gate oxide leakage current.

D. Datta et al. / Microelectronics Journal 37 (2006) 537–545 545

Based on the potential profile, we have achieved the profile

of bulk as well as inversion charges based on symmetric

linearization method. Symmetric linearization also helps us

to reach the essential symmetry with respect to source/drain

interchange, which is a necessary feature of symmetric

DG MOSFETs. In NQS profile, we have derived the

variation of inversion charge with respect to time, where

gate voltage changes with time in a ramp profile or, in 1D

gaussian profile. We have also investigated the expected

Gummel Symmetry in NQS profile. In the present paper, we

have not included the quantum mechanical space quantiza-

tion effect, but the symmetricity of the charge-sheet model

gives the expected behavior of the bulk and inversion charge

density through out the device. This symmetrically

linearized model helps us to derive the transcapacitance

and transconductance profile of both symmetric and

asymmetric DGMOSFET. Finally, scaling limit projections

indicate that individual DG MOSFETs with good turn-off

behavior are feasible at 10 nm scale; however, practical

exploitation of these devices towards giga-scale integrated

systems requires significant improvement in process

control. This model also highlights the device physics

essential to predict the transport phenomena in

DG MOSFETs. The numerical model presented here is

used as tool for simulating modern thin gate DG MOSFETs

used in VLSI circuits.

Acknowledgements

The authors would like to acknowledge the financial

support provided by Ministry of Human Resources and

Development, Govt. of India under Thrust Areas of

Technical Education scheme (No. F27-3/2002TS.V)

References

[1] Y. Taur, C.H. Wann, D. Frank, 25 nm CMOS design considerations,

IEDM Tech. Digest (1998) 789–792.

[2] International Technology Roadmap of Semiconductors, Update, Free

available athttp://www.itrs.net/common/2004update 2004.

[3] D.J. Frank, S. Laux, M. Fischetti, Monte carlo simulation of a 30 nm

dual-gate MOSFET: how short can Si go?, IEDM Tech. Digest (1992)

553–556.

[4] Chen Wanqiang, F.Register Leonard, K.Banerjee Sanjay, Scattering

in a nano-scaleMOSFET: a quantum transport analysis, IEDM (2003).

[5] Brian Winstead, Umberto Ravaioli, A quantum correction based on

schrodinger equation applied to monte carlo device simulation, IEEE

Trans. Electron Dev. 50 (2) (2003) 440–446.

[6] J.A. Lopez-Villanueva, P. Cartujo-Cassinello, F. Gamiz, J. Banqueri,

A.J. Palma, Effects of the inversion layer centroid on the performance

of DG MOSFETs, IEEE Trans. Electron Dev. 47 (1) (2000) 141–146.

[7] Z. Ren, M. Lundstrom, Extraordinarily high drive currents in

asymmetrical DG MOSFETs, Superlattices Microstruct. (2000).

[8] A.R. Boothroyed, S.W Tarasewicz, C. Slabby, MISNAN -A

physically based constant MOSFET model for CAD applications,

IEEE Trans. Electron Dev. 10 (1991) 1512–1529.

[9] G. Baccarani, S. Reggiani, A compact double-gate MOSFET model

comprising quantum-mechanical and nonstatic effects, IEEE Trans.

Electron Dev. 8 (1999) 1656–1666.

[10] Y. Liu, et al. Ultra thin channel vertical DG MOSFET fabricated by

using ion-bombardment-retarded etching, IEEE Trans. on Electron

Devices, Accepted for Future Publication.

[11] Man Wong, Xuejie Shi, On the threshold voltage of symmetrical DG

MOS capacitor with intrinsic silicon body, IEEE Trans. Electron Dev.

51 (2004) 1600–1604.

[12] M. Miura-Mattausch, W. Feldmann, A.R. Ahm, M. Bollu,

D. Savignac, Unified complete MOSFET model for analysis of digital

and analog circuits, IEEE Trans. Comput-Aid Design 15 (1996) 1–7.

[13] K. Joardar, K. Joardar, K.K. Gullapalli, C.C. McAndrew,

M.E. Burnham, A. Wild, An improved MOSFET model for circuit

simulation, IEEE Trans. Electron Dev. 45 (1) (1998) 134–148.

[14] Xiaoping Liang, Y. Taur, A 2-D analytical solution for SCEs in DG

MOSFETs, IEEE Trans. Electron Dev. 51 (2004) 1385–1391.

[15] J.R. Brews, A charge sheet model for MOSFET, Solid State Electron

21 (1978) 345–355.

[16] D.E. Ward, R.W. Dutton, A charge oriented model for MOS transistor

capacitances, IEEE J. Solid State circuits SC-13 (1978) 703–708.

[17] S. Dasgupta, P. Chakrabarti, Effect of ionising radiation on the

characteristics of a MOSFET, IEE Circ. Dev. Syst. 147 (2000) 133–

138.

[18] T.L. Chen, G. Gildenblat, Symmetric bulk charge linearization in

charge-sheet MOSFET model, Electron Lett. 37 (2001) 791–793.

[19] C.C. Enz, MOS transistor modeling dedicated to low current and low

voltage analog circuit design and simulation Presented at Sixth

Brazilian School of Microelectronics, May 1996.

[20] Cunha, et al., An explicit physical model for long channel MOS

transistor including small signal parameters, Solid State Electron 38

(1995) 1945–1952.