Vous êtes sur la page 1sur 24

Applied Numerical Mathematics 43 (2002) 229252

www.elsevier.com/locate/apnum
A relaxation scheme for the hydrodynamic equations
for semiconductors
Ansgar Jngel
a,
, Shaoqiang Tang
a,b
a
Fachbereich Mathematik und Statistik, Universitt Konstanz, 78457 Konstanz, Germany
b
Department of Mechanics and Engineering Science, Peking University, Beijing 100871, Peoples Republic of China
Abstract
In this paper, we shall study numerically the hydrodynamic model for semiconductor devices, particularly in
a one-dimensional n
+
nn
+
diode. By using a relaxation scheme, we explore the effects of various parameters,
such as the low eld mobility, device length, and lattice temperature. The effect of different types of boundary
conditions is discussed. We also establish numerically the asymptotic limits of the hydrodynamic model towards
the energy-transport and drift-diffusion models. This veries the theoretical results in the literature.
2002 IMACS. Published by Elsevier Science B.V. All rights reserved.
1. Introduction
One of the main goals in semiconductor device modeling is to establish a hierarchy of models,
which allows for the choice of appropriate models for specic semiconductor applications [25].
Monte Carlo simulations of the kinetic Boltzmann equation provide a very accurate description of
charge transport in submicron devices [15]. However, their use is not practical for computer aided
design because of the large computer times needed. Macroscopic models derived from the Boltzmann
equation seem to be a compromise between physical accuracy and computational effort. The main
classes of macroscopic semiconductor models are the drift-diffusion, energy-transport and hydrodynamic
equations [29,36].
The drift-diffusion models which are the most popular ones were rst proposed in 1950 by Van
Roosbroeck [33]. The energy-transport equations include also the carrier energy (or temperature)
which is constant in the drift-diffusion equations, and have been suggested one decade later by
Stratton [39]. The drift-diffusion and energy-transport models can be formally derived by a Chapman
*
Corresponding author.
E-mail addresses: juengel@fmi.uni-konstanz.de (A. Jngel), tangs@fmi.uni-konstanz.de (S. Tang).
0168-9274/02/$ see front matter 2002 IMACS. Published by Elsevier Science B.V. All rights reserved.
PII: S0168- 9274( 01) 00182- 9
230 A. Jngel, S. Tang / Applied Numerical Mathematics 43 (2002) 229252
Enskog type expansion method from the Boltzmann equation [7,8]. The drift-diffusion equations give
satisfactory results for semiconductor devices with a typical size of a few microns and moderately
applied voltage [35], whereas energy-transport models can also be used for certain submicron
devices [14].
The hydrodynamic equations have been introduced by Bltekjr [9] and subsequently thoroughly
investigated by Baccarani and Wordeman [6]. They can be derived from the Boltzmann equation by
using a moment method. This yields usually a set of equations for the carrier density, momentum and
energy which is not in closed form. To obtain a closed set of equations, often the Fourier law for the
heat ux is taken [9]. For different approaches of the derivation of the hydrodynamic equations and a
discussion of the closing problem, we refer to [3].
The hydrodynamic equations derived by Bltekjr, Baccarani and Wordeman read as follows:
n
t

1
q
divJ =0, (1)
J
t

1
q
div
_
J J
n
_

qk
B
m

(nT ) +
q
2
m

nV =C
J
, (2)
E
t
div
_
m

2q
3
J|J|
2
n
2
+
5
2
k
B
q
T J +T
_
=J V +C
E
, (3)

s
V =q(n C). (4)
Here, the physical variables are the electron density n, the current density J, the energy density E,
and the electrostatic potential V. The constants are the elementary charge q, the Boltzmann constant k
B
,
the effective electron mass m

, and the permittivity constant


s
. The doping concentration characterizing
the device under consideration is denoted by C =C(x). We assume the following constitutive relations.
The energy density is given as the sum of kinetic and thermal energy
E =
m

2q
2
|J|
2
n
+
3
2
k
B
T n =
1
2
m

n|u|
2
+
3
2
k
B
T n,
where the electron velocity u is dened by J = qnu. The momentum and energy relaxation terms,
respectively, are
C
J
=
J

p
,
C
E
=
1

w
_
m

2q
2
|J|
2
n
+
3
2
k
B
(T T
0
)n
_
,
where T
0
is the lattice temperature,

p
=
po
_
T
T
0
_
r
,
w
=
wo
T
T +T
0
+
1
2

p
are the momentum and energy relaxation times, respectively, with

po
=
m

n
q
,
wo
=
3
n
k
B
T
0
2qv
2
s
, (5)
A. Jngel, S. Tang / Applied Numerical Mathematics 43 (2002) 229252 231

n
is the low-eld mobility and v
s
the saturating velocity. Finally, the heat conductivity is assumed to be
=
_
5
2
+c
_
k
2
B

n
q
nT
_
T
T
0
_
r
,
where c, r R are some phenomenological constants. In the numerical simulations we use c =r =1.
Eqs. (1)(4) have to be solved in a bounded domain. In this paper we will present numerical
simulations for the above equations in one space dimension. We use homogeneous Neumann boundary
conditions for n, u and T at the domain boundary (see Section 4.2 for Dirichlet boundary conditions for
n and T ). Moreover, we impose initial conditions for n, u and T (see Section 3.1).
In physical situations where the mean free path of the particles is much smaller than the typical device
length and the momentum relaxation time constant
po
is much smaller than the energy relaxation time
constant
wo
, the hydrodynamic equations reduce in the relaxation-time limit formally to the energy-
transport equations (see Section 2). This limit has been proved rigorously in [17] under the assumption
of uniform L

bounds. If the momentum and energy relaxation times are of the same order, but the mean
free path is much smaller than the typical device length, we obtain formally the drift-diffusion equations
from the hydrodynamic model. This limit has been shown rigorously for constant temperature in [23].
The asymptotic limit in the full model has been studied under some conditions in [1,12]. For an overview
of these limits, see [24].
The main objectives of this paper are rst to adopt the relaxation scheme [22] in order to solve
numerically the hydrodynamic model in one space dimension and secondly, to perform the above
asymptotic limits numerically. This allows to determine the numerical values for which the solution
to the hydrodynamic model behaves like the solution of the drift-diffusion or energy-transport equations.
Moreover, we will illustrate the effects of the mobility constant, lattice temperature and channel length
of a simulated n
+
nn
+
diode.
A relaxation model was rst rigorously studied in [28], and a relaxation scheme was proposed in [22].
Various generalizations have then been made, e.g., discrete BGK schemes [4,30]. The basic idea is as
follows. In general, a set of conservation laws, usually quasilinear, may be derived as a macroscopic
model from a Boltzmann type equation with certain equilibrium states (e.g., local Maxwellians). This
Boltzmann type equation is semilinear, yet contains an additional variable, namely the momentum.
It is therefore much more expensive to simulate numerically. However, we may design a discrete
BGK equation instead, i.e., an articial Boltzmann equation with nite discrete moments and suitable
Maxwellians, which are constructed in such a way to give the desired set of hyperbolic conservation laws
when performing the limiting process. A relaxation model in [22] is a special case when we take only
two velocities as moments.
To solve this relaxation model, we may apply a splitting method, i.e., rst solving an ODE step and
then solving a linear convection step. As the relaxation parameter tends to zero, this solution tends to the
solution of the original problem (see Section 3). The resulting scheme has the advantage that it possesses
a modular structure, which is particularly good for coding and for higher space dimensions. An extensive
exploration on the stability, efciency, as well as accuracy has been made for general discrete BGK
models in [4]. It is also shown to be robust when applied to hyperbolicelliptic systems, and strongly
degenerate parabolic systems in one dimension and multi-dimensions [5,31]. We thus deem it suitable
to treat semiconductor devices, where various complexities are present, such as being of hyperbolic
parabolic type, having stiff source terms, etc.
232 A. Jngel, S. Tang / Applied Numerical Mathematics 43 (2002) 229252
The numerical discretization of the transient full hydrodynamic model in two space dimensions using
a discrete BGK method will be presented in a forthcoming publication [26].
There are several other techniques to discretize the hydrodynamic equations. First the Scharfetter
Gummel method has been generalized for these equations, in particular for subsonic ow [34]. Later,
second-order upwind shock-capturing methods have been used for transonic ow [16]. In recent years,
numerical techniques like streamline-diffusion schemes [20], nite element methods of the RungeKutta
discontinuous Galerkin scheme [13], nite difference methods of the ENO (essentially non-oscillatory)
scheme [19,37], and UNO (uniformly non-oscillatory) schemes with the NessyahuTadmor method [3,
32] have been developed.
This paper is organized as follows. In Section 2 we scale Eqs. (1)(4) appropriately and explain
the relaxation-time limits towards the energy-transport and the drift-diffusion equations in more detail.
Section 3 is concerned with the numerical discretization of Eqs. (1)(4) in one space dimension.
Numerical simulations for a stationary one-dimensional n
+
nn
+
diode which can be considered as a
benchmark problem are presented in Section 4.
2. Scaling of the equations and asymptotic models
In this section we scale Eqs. (1)(4) appropriately and derive the energy-transport and drift-diffusion
models by means of formal asymptotic analysis.
2.1. Scaling
We introduce the thermal voltage U
T
=k
B
T
0
/q, the mean free path
=
po
_
k
B
T
0
m

,
and the scaled Debye length
=
_

s
U
T
qC
m
L
2
,
where C
m
is a typical doping concentration and L is a typical device length. Furthermore, we dene the
dimensionless parameters
=

L
, =
_

po

wo
. (6)
Then, with the scaling
n C
m
n, t
po
t, C C
m
C, x Lx,
V U
T
V, T T
0
T, J
_
q
2
U
T
C
m

po
/Lm

_
J,
we obtain the scaled equations
n
t
divJ =0, (7)
A. Jngel, S. Tang / Applied Numerical Mathematics 43 (2002) 229252 233
J
t

2
div
_
J J
n
_
(nT ) +nV =
J
T
r
, (8)
_
|J|
2
2n
+
3
2
nT

2
_
t
div
_

2
J|J|
2
2n
2
+
5
2
T J + T
_
=J V
2
_
T
T +1
+

2
2
T
r
_
1
_
|J|
2
2n
+
3
2
n(T 1)

2
_
, (9)

2
V =n C. (10)
We used the same notation for the scaled and unscaled variables. In Eq. (9) we have set
=
_
5
2
+c
_
T
1+r
n.
2.2. Asymptotic models
In order to obtain the energy-transport and drift-diffusion models, we rescale Eqs. (7)(9) by t t /
2
:
n
t
divJ =0, (11)

2
J
t
+
2
div
_
J J
n
_
(nT ) +nV =
J
T
r
, (12)
_

2
|J|
2
2n
+
3
2
nT
_
t
div
_

2
J|J|
2
2n
2
+
5
2
JT + T
_
=J V
_
T
T +1
+

2
2
T
r
_
1
_

2
|J|
2
2n
+
3
2

2
n(T 1)
_
. (13)
The energy-transport equations are obtained by assuming that
1, 1.
The relation 1 holds if the kinetic energy associated with the velocity needed to cross the device in
time
po
is very large compared with the thermal energy. The relation 1 also means that we study the
system at large times of the order of 1/
2
. Furthermore, it holds 1 if the kinetic energy associated
with the saturating velocity is much smaller than the thermal energy.
We formally perform the limit 0, 0 such that /

0
, where
0
> 0 is some constant.
The limit /

0
means that the velocity needed to cross the device in time
po
is assumed to be of
the same order as the saturating velocity. We obtain the equations:
n
t
divJ =0, (14)
J =T
r
_
(nT ) nV
_
, (15)
_
3
2
nT
_
t
div
_
5
2
T J + T
_
=J V
3
2
n(T 1)
(T )
, (16)

2
V =n C, (17)
where (T ) =
0
T/(T +1). Notice that this energy-transport model is not of the general form derived
in [14] expect for c =r =0.
234 A. Jngel, S. Tang / Applied Numerical Mathematics 43 (2002) 229252
For the derivation of the drift-diffusion model, we x the parameter > 0 and let formally 0 in
Eqs. (11)(13) to obtain T =1 and
n
t
divJ =0, (18)
J =n nV, (19)

2
V =n C. (20)
3. Numerical scheme
In this section, we shall put the model into a more concise form. Then we shall describe our numerical
scheme in three parts, namely, the overall second-order splitting method, the relaxation scheme for the
convection, and the treatment of the boundaries and diffusion.
3.1. Reformulation of the model
We recast the hydrodynamic model in one space dimension as follows.
U
t
+A(U)
x
=
_
B(U, U
x
)
_
x
+S(U), (21)
coupled with the Poisson equation

s
V
xx
=q
_
n C(x)
_
. (22)
The vector quantities are
U =
_
n

E
_
, A(U) =
_
_
_
_
_
_

2
3
_

2
n
+
E
m

_
5E
3n

3
3n
2
_

_
,
B(U, U
x
) =
_
0
0
T
x
_
, S(U) =
_
_
_
_
_
0

p
+
nqV
x
m

qV
x

E E
0

w
_

_
.
Here = nu = J/q, E =
1
2
m

nu
2
+
3
2
nk
B
T, and E
0
=
3
2
nk
B
T
0
is the rest energy density. We recall
the relaxation coefcients and heat diffusion coefcient

p
=
m

n
q
_
T
T
0
_
r
, (23)

w
=
1
2

p
+
3
n
k
B
T
0
T
2qv
2
s
(T +T
0
)
, (24)
=
_
5
2
+c
_
nk
2
B

n
T
q
_
T
T
0
_
r
, (25)
where we use c =r =1.
A. Jngel, S. Tang / Applied Numerical Mathematics 43 (2002) 229252 235
Table 1
Physical parameters
Parameter Physical meaning Numerical value
q elementary charge 1.6 10
19
C
m

effective electron mass 0.26 9.11 10


31
kg

s
permittivity constant 11.7 8.85 10
12
F/m

n
low eld mobility constant 0.1 m
2
/Vs
k
B
Boltzmann constant 1.38 10
23
J/K
T
0
lattice temperature 300 K
n
i
intrinsic electron concentration 1.4 10
16
m
3
v
s
saturating velocity 1.03 10
5
m/s
We list the physical parameters used in our simulations in Table 1.
The initial and boundary conditions are assigned as:
n(x, 0) =C(x), (26)
u(x, 0) =0, (27)
T (x, 0) =T
0
, (28)
n
x
(0, t ) =n
x
(L, t ) =0, (29)
u
x
(0, t ) =u
x
(L, t ) =0, (30)
T
x
(0, t ) =T
x
(L, t ) =0. (31)
Taking into account the conservation of electrons, we notice that the reecting boundary condition on
n implies the xed boundary condition
n(0, t ) =C(0), n(L, t ) =C(L).
For the electric eld with an applied voltage V
b
, the boundary conditions are
V(0) =
T
0
q
ln
_
n(0, t )
n
i
_
, (32)
V(L) =
T
0
q
ln
_
n(L, t )
n
i
_
+V
b
. (33)
3.2. Second-order RungeKutta splitting scheme
The system (21) can be solved by a splitting method. Given some data at the kth time step t =t
k
, one
rst solves an ODE step, namely the initial-value problem with electric eld V
k
x
corresponding to the
electron concentration n(x, t
k
),
U
t
=
_
B(U, U
x
)
_
x
+S(U), U
_
x, t
k
_
=U
k
(x). (34)
Denote the solution at t =t
k
+t as U
k+1/2
=R(U
k
, V
k
x
, t ). As we shall explain in a later subsection,
the diffusion term (B(U, U
x
))
x
is expressed explicitly by quantities at neighboring grid points. This
makes (34) an ODE system.
236 A. Jngel, S. Tang / Applied Numerical Mathematics 43 (2002) 229252
Secondly, one solves a convection step, namely the initial-boundary-value problem for the homoge-
neous system
U
t
+A(U)
x
=0, U
_
x, t
k
_
=U
k+1/2
, U
x
(0, t ) =U
x
(L, t ) =0, (35)
by a hyperbolic problem solver. The solution after one time step is denoted as U
k+1
=W(U
k+1/2
, V
k
x
, t )
(see Section 3.3 for an explicit formula).
Finally the electric eld is updated from the Poisson equation (22) with electron concentration
n
k+1
=rst component of U
k+1
, denoted as V
k+1
x
=P(n
k+1
).
This is just a rst-order splitting. There are different ways to make it of high-order accuracy, e.g., [10,
38]. We simply apply the mid-point method to combine two Euler-forward steps for second-order
accuracy. More precisely, the scheme is:
(1) Set the initial data U
0
=U(x, 0) and update the electric eld V
0
x
=P(n
0
).
(2) For k =0, . . . , k
T
1, compute U
k+1
, V
k+1
x
as follows:
(a) Set (U
[0]
, V
[0]
x
) =(U
k
, V
k
x
).
(b) Compute (U
[1]
, V
[1]
x
) and U
[2]
as follows:
U
[1]
=
_
n
[1]
,
[1]
, E
[1]
_
=U
[0]
+
_
W
_
R
_
U
[0]
, V
[0]
x
, t
_
, V
[0]
x
, t
_
U
[0]
_
/2,
V
[1]
x
=P
_
n
[1]
_
,
U
[2]
=W
_
R
_
U
[1]
, V
[1]
x
, t
_
, V
[1]
x
, t
_
.
(c) Set U
k+1
=(U
[1]
+U
[2]
)/2, V
k+1
x
=P(n
k+1
).
We note that this is an explicit splitting, hence one should take t (x)
2
for stability. It can be
improved by applying implicit techniques.
3.3. Relaxation scheme for the convection step
The numerical resolution for homogeneous hyperbolic systems has been the main advance in
computing science during the last two decades. We adopt a relaxation scheme in the simulations. That is,
we approximate the quasi-linear system (35) by a semi-linear one,
U

t
+Y

x
=0, (36)
Y

t
+
2
U

x
=
A(U

) Y

, (37)
where is a small parameter and is a constant, larger than the maximum wave speed of the original
system (sub-characteristic condition) for stability. As approaches towards 0, formally Y

approaches
towards A(U

). In turn the rst equation approximates (35). Rigorous results are obtained, e.g., in [28,
30] and references therein.
At the numerical level, this relaxation scheme bears many nice features, such as high accuracy,
modular structure, easy to code, ready for high-dimensional generalization, etc. Numerically one may
use the same splitting method in the previous section for this system. Since (37) is semilinear, the linear
convection part (for (36), (37) without the term on the right-hand side) is readily solved by some existing
scheme, e.g., a second-order MUSCL type scheme with minmod limiter (see, e.g., [27]). In the ODE step
A. Jngel, S. Tang / Applied Numerical Mathematics 43 (2002) 229252 237
(for (36), (37) without the terms Y

x
and
2
U

x
), on the other hand, it is observed that U

keeps unchanged,
therefore the Y

-equation can be solved explicitly as


Y

(t ) =Y

(0) exp
_

_
+A
_
U

_
_
1 exp
_

__
.
A relaxed scheme is obtained, at this step, when putting to zero, or projecting Y

to the ux
function (Maxwellian) A(U

). Note that then becomes the only numerical parameter introduced by


the relaxation method. We introduce U
k
i
= U
k
(x
i
) and = t /x. The explicit expression of the
relaxed scheme for one-step marching U
k+1
=W(U
k+1/2
, V
k
x
, t ) is
U
k+1
i
= (1 )U
k+1/2
i
+
_
M
+
i1
+M

i+1
_

(1 )
2

_
minmod
_
M
+
i+1
M
+
i
, M
+
i
M
+
i1
_
minmod
_
M
+
i
M
+
i1
, M
+
i1
M
+
i2
_
+minmod
_
M

i+2
M

i+1
, M

i+1
M

i
_
minmod
_
M

i+1
M

i
, M

i
M

i1
__
with
M

i
=
U
k+1/2
i
A(U
k+1/2
i
)/
2
.
3.4. Source terms and boundaries
In the ODE step, the rst equation in (34) implies that n remains unchanged, and consequently so for
the electric eld. However, the relaxation times
p
,
w
and the heat diffusion parameter are expressed in
terms of the temperature T , and therefore of the energy E. The ODE system cannot be solved explicitly
in general. It takes the following form:

t
=

p
+
nqV
x
m

, (38)
E
t
=qV
x

E E
0

w
+(T
x
)
x
. (39)
We may apply a RungeKutta method to integrate this system. As a remark, since t (x)
2
, even
the forward Euler method is enough to maintain the accuracy. If an implicit second order splitting is
employed, one may use a second-order integrator for this ODE step to match the accuracy.
It is worth mentioning the treatment of the heat diffusion term. In our code, it is approximated in
conservative form with second-order accuracy by
(T
x
)
x

1
2(x)
2
_
(
i+1
+
i
)(T
i+1
T
i
) (
i
+
i1
)(T
i
T
i1
)
_
, (40)
where
i
is the local heat diffusion parameter in (25) with n
i
, T
i
plugged in.
The Poisson equation (22) is discretized by the standard central difference scheme. The resulting tri-
diagonal algebraic system can be readily solved.
In all the discretizations, reecting boundary conditions are used for n, u, T , and V
x
, i.e., with
reectingly-valued ghost-points once needed. This is consistent with the boundary conditions in the
continuous form (26)(31).
238 A. Jngel, S. Tang / Applied Numerical Mathematics 43 (2002) 229252
4. Numerical simulations
In this section, we shall rst describe the numerical tests with our scheme. Then we shall illustrate
the effects of the mobility constant, channel length, and lattice temperature. We also demonstrate the
effects of taking Dirichlet boundary conditions for n and T , instead of the aforementioned Neumann
boundary conditions. Finally we shall explore the numerical limits of the hydrodynamic model towards
the dimensionalized energy-transport model and drift-diffusion model, respectively.
4.1. Benchmarks on the numerical scheme
We make numerical tests on an n
+
nn
+
ballistic silicon diode with an applied voltage of 1.5 V. The
semiconductor domain is described by the interval [0, L] with L =0.6 m. The channel length is 0.4 m.
The doping prole is:
C(x) =
_
2 10
21
m
3
, x (0.1 m, 0.5 m),
5 10
23
m
3
, elsewhere.
(41)
We make a convergence study with different meshes. In all these tests, the numerical relaxation
parameter in the relaxation scheme is set as = 2.5. A computation shows that this value satises the
subcharacteristic condition. We perform simulations with successively double space grid numbers, i.e.,
with 100 grid points (x =6 10
3
m), 200 grid points (x =3 10
3
m), 400 grid points (x =
1.5 10
3
m), 800 grid points (x =7.5 10
4
m), and 1600 grid points (x =3.75 10
4
m),
respectively.
The numerical solution with 200 gridpoints is displayed in Fig. 1. In the rst few picoseconds,
oscillations occur due to the sudden application of the electric eld and the initial conditions. They
are damped out gradually, and the solution tends to a steady state. In fact, it is fairly stationary-like at
t = 5 ps. The units are taken as 10
1
m for the space variable x, 10
21
m
3
for the concentration n,
10
5
m/s for the velocity u, 10
1
Jm
3
for the energy density E, 10
1
eV = 1.6 10
20
J for the total
energy w = E/n, 10
3
K for the temperature T , and 10
6
V/m for the electric eld V
x
. This scale of
units will be used throughout the gures hereafter.
We depict the stationary solution at t = 15 ps in Fig. 2. In the velocity prole, it is observed that
besides an overshoot in the second junction, an even bigger hump appears in a fairly wide region around
the rst junction. This hump has already been observed in the simulation of the hydrodynamic model [11].
It results from our choice of the parameters. The change of the heat diffusion constant exponent r may
give different proles [19]. The low eld mobility
n
also makes big differences, e.g., as reported for
a GaAs diode [11]. We shall describe the numerical simulations with different constants
n
in the next
subsection. We also observe a cooling zone around the rst junction, and a heating zone around the
second with highest temperature about 5 times the lattice temperature.
We now describe the numerical convergence study. The solutions with different meshes are shown
in Fig. 3. For a better presentation of the differences in the concentration and energy density proles,
logarithmic plots are used. With ner grids, the velocity overshoot in the second junction clearly becomes
sharper. It is likewise in the total energy prole. These are known effects of the numerical viscosity. For
ner mesh, numerical viscosity is smaller, thus the numerical solution has less smearing around the
discontinuities or the place where large gradients occur. Differences in other quantities are relatively
small. For the electric eld, the difference is even negligible. This can be explained by analyzing
A. Jngel, S. Tang / Applied Numerical Mathematics 43 (2002) 229252 239
Fig. 1. Numerical solution with 200 grid points (x =3 10
3
m), (t , x) [0, 10] [0, 6].
the steady states. As the numerical current should keep constant, the difference in u is reected in n
only reciprocally. Because n ranges from around 10 to 5000, this difference is barely observable. The
difference in n interferes the electric eld through the Poisson equation. The twice integration therefore
further diminishes the difference.
Let us make a quantitative analysis of the differences. Taking the solution with the nest grid
(1600 grid points) as an exact solution, we list the L

and L
1
errors in Table 2. The numerical
convergence rates for the L
1
error are also computed. The units of the L

errors are the same as the


240 A. Jngel, S. Tang / Applied Numerical Mathematics 43 (2002) 229252
Fig. 2. Numerical solution with 1600 grid points (x =3.75 10
4
m).
corresponding quantities, respectively, whereas the units for L
1
errors should take the unit of space into
account.
As explained before, large gradients (discontinuities) occurs in the solution around the second junction.
Around this point, our scheme only maintains rst order accuracy, which is the case for most existing
hyperbolic solvers. Moreover, the L

error is reached here. Different meshes yield different smearing


effects, and right at the spike, the L

difference is not negligible. Away from this point, the L

difference is indeed very small. We also remark that the spike may be involved with a numerical
artifact similar to that in a slowly moving discontinuity [21]. The losing of accuracy around the
discontinuity makes the L
1
convergence rate between 1 and 2. For a more comprehensive description,
see Fig. 4.
4.2. The effects of the mobility constant, channel length, lattice temperature, and Dirichlet boundary
conditions
It is known that the solution of the hydrodynamic model changes along with the physical parameters.
In particular, we display the numerical results for different (constant)
n
in Fig. 5. When decreasing
A. Jngel, S. Tang / Applied Numerical Mathematics 43 (2002) 229252 241
Fig. 3. Numerical solutions with successive double grid points. Dotted: 100 grid points; dashed: 200 grid points; solid: 400 grid
points; heavy-dotted: 800 grid points.
the mobility constant
n
, it turns out that there are not much change in the concentration n, nor in the
electric eld V
x
. A distinct feature lies in the gradual diminishing of the hump in the velocity prole.
A at interval appears within the channel for mobility
n
0.04 m
2
/Vs. Moreover, the temperature
prole gets atter, yet keeps slightly higher than the lattice temperature T
0
= 300 K on the right
boundary.
1
Secondly, if we simulate a device with shorter channel length, the basic picture is quite similar. For
instance, Fig. 6 depicts the solution at t =15 ps for a device with a 0.2 m channel. For
n
=0.1 m
2
/Vs,
the velocity hump is more profound than in the previous case. An energy density peak appears after the
second junction. The heating effect in the second junction is even stronger, with highest temperature
about eight times the lattice temperature. As there is a longer n
+
range, the temperature on the right
boundary almost cools down to the lattice temperature. The electric eld intensies and gets sharper
1
As the mobility
n
becomes smaller, also the parameter is becoming smaller (see (5) and (6)), whereas is xed. This
will therefore correspond to the drift-diffusion limit as explained in Section 2.2. By Eq. (13), we expect that the temperature T
becomes closer to the lattice temperature. A numerical limit study will be presented in Section 4.3.
242 A. Jngel, S. Tang / Applied Numerical Mathematics 43 (2002) 229252
Table 2
Numerical errors for different number of grid points
Grid number Concentration n Momentum =nu
L

error L
1
error rate L

error L
1
error rate
100 52.398 21.566 1.67 32.636 21.411 1.92
200 46.019 6.7560 1.31 14.902 5.6418 1.79
400 20.522 2.7227 0.50 7.7074 1.6296 2.02
800 19.797 1.9311 6.8576 0.4030
Grid number Energy E Electric eld V
x
L

error L
1
error rate L

error L
1
error rate
100 64.681 52.995 1.32 0.9787 0.5162 2.14
200 46.911 21.203 2.05 0.4713 0.1173 0.88
400 19.840 5.1181 0.89 0.1583 0.0637 0.76
800 16.906 2.7647 0.1044 0.0377
Grid number Velocity u Temperature T
L

error L
1
error rate L

error L
1
error rate
100 1.2121 0.2987 1.22 0.2420 0.2224 1.51
200 1.2257 0.1278 1.93 0.1675 0.0782 1.66
400 0.8105 0.0336 1.07 0.0614 0.0248 1.58
800 0.6609 0.0160 0.0376 0.0083
around the junctions. When the low eld mobility decreases, the velocity hump again decreases in the
amplitude, and almost disappears at around
n
0.025 m
2
/Vs. The energy density, total energy, and
temperature within the channel decrease considerably. This may be explained as above from the change
of the parameters and .
Thirdly, we simulate a Si-diode with 50 nm-channel. Under different model equations, this kind of
device has been studied in [2]. We consider a 250 nm long device with doping prole
C(x) =
_
2 10
21
m
3
, x (0.1 m, 0.15 m),
5 10
23
m
3
, elsewhere.
(42)
A voltage of 0.6 V is applied. The other physical constants remain the same as before. As the system
tends to equilibrium quicker at this length scale, we illustrate the solution at time t = 5 ps in Fig. 7.
The concentration differs much from the doping prole, particularly by a shift to the right. This can be
explained by the positive mean velocity eld. For all the mobility constants we have tried, no at velocity
interval is observed. Moreover, the hottest position is located quite far away from the second junction,
moving gradually to the left when the mobility constant decreases. It is similar for the maximal velocity
point. Quite interestingly, we observe that the cooling zone moves to the right to such an extent that
it is relatively cool inside the channel.
2
Though the right n
+
region is twice the channel length, it is
2
This means that the total energy goes completely in the kinetic energy.
A. Jngel, S. Tang / Applied Numerical Mathematics 43 (2002) 229252 243
Fig. 4. Convergence study of the scheme. Left: L

errors; right: L
1
errors. Legend: +: concentration n; : momentum ;
: energy density E; : electric eld V
x
; o: velocity u; : temperature T .
not enough to cool the device down to the lattice temperature. These facts together yield energy density
proles lifted on the right part.
3
Now we investigate the effect of the lattice temperature. Solutions at T
0
=300 K, 200 K, 100 K, 80 K,
60 K are displayed in Fig. 8. Although this is by no means a limit study for T
0
0, the numerical
results suggest that n, u, E, w, and V
x
might converge to certain proles, yet not clear for the scaled
temperature T/T
0
. We observe a kink developing around x 0.35 m in the proles of u, E, w, and a
hump in n. A second temperature peak lies at this point. For T
0
200 K, this temperature peak surpasses
the hot point near the second junction.
4
3
Notice that the parameter becomes larger as the channel length decreases, i.e., the system is far away from the energy-
transport regime and convective effects are important.
4
As T
0
decreases, the parameter becomes smaller. However, we are not in the drift-diffusion regime, since becomes
larger as T
0
0.
244 A. Jngel, S. Tang / Applied Numerical Mathematics 43 (2002) 229252
Fig. 5. Numerical solutions for different low eld mobility. Dotted:
n
= 0.1 m
2
/Vs, dash-dot:
n
= 0.08 m
2
/Vs, dashed:

n
=0.06 m
2
/Vs, solid:
n
=0.05 m
2
/Vs, heavy-dotted:
n
=0.04 m
2
/Vs.
Finally, instead of the Neumann boundary conditions (29) and (31) on n and T , we impose xed
boundary conditions
n(0, t ) =C(0), (43)
n(L, t ) =C(L), (44)
T (0, t ) =T
0
, (45)
T (L, t ) =T
0
. (46)
As explained before, the Neumann boundary conditions on n are equivalent to the Dirichlet boundary
conditions. However, this is not the case for T . In fact, we observed before an elevation of temperature
on the right boundary from the lattice temperature T
0
. Quite naturally, the Dirichlet boundary conditions
lays a stronger effect when the solution under Neumann boundary conditions differs more from T
0
at
the boundary. In Fig. 9, we compare the solutions with the same parameters as those in Section 4.1.
The deviation is barely observable. However, for the 50 nm-channel, the difference is fairly big. See
Fig. 10.
A. Jngel, S. Tang / Applied Numerical Mathematics 43 (2002) 229252 245
Fig. 6. Numerical solutions for different low eld mobility for channel length 0.2 m. Dotted:
n
= 0.1 m
2
/Vs, dash-dot:

n
=0.06 m
2
/Vs, dashed:
n
=0.03 m
2
/Vs, solid:
n
=0.025 m
2
/Vs, heavy-dotted:
n
=0.02 m
2
/Vs.
4.3. Approximation to the energy-transport model and the drift-diffusion model
As discussed in Section 2, the hydrodynamic model approaches to the energy-transport model in
the limit 0, 0 with the ratio / xed. In our numerical study, we x all other parameters
as in Section 4.1 but
n
=
n0
, v
s
= v
s0
, with
n0
= 0.1 m
2
/Vs and v
s0
= 1.03 10
5
m/s, and
is a parameter. Correspondingly, we have =
0
, =
0
with
0
=
n0

k
B
T
0
/qL = 0.10315,

0
=
_
2m

v
2
s0
/3k
B
T
0
=2.0116.
The limiting system as 0 is the energy-transport model (14)(17) in non-dimensionalised form.
In Fig. 11 we illustrate the approximation. Taking the same doping prole as before, we solve (21)
numerically for =1, 0.5, 0.2, 0.1, and 0.05, respectively. The solutions are plotted at time T =15 ps/.
The numerical solutions seem to converge. A spike appears again in the prole of the rescaled velocity
u/. We note that the time step size and the discrete BGK parameter are adjusted in the simulations, for
the sake of stability and computing load. Compared with the stationary solution of the energy-transport
model (solved with the numerical method of [18]) in Fig. 12, the numerical limit of the hydrodynamic
246 A. Jngel, S. Tang / Applied Numerical Mathematics 43 (2002) 229252
Fig. 7. Numerical solutions for different low eld mobility for channel length 50 nm. Dotted:
n
= 0.1 m
2
/Vs, dash-dot:

n
=0.08 m
2
/Vs, dashed:
n
=0.06 m
2
/Vs, solid:
n
=0.04 m
2
/Vs, heavy-dotted:
n
=0.02 m
2
/Vs.
model agrees reasonably well, particularly the density prole and the electric eld. The deviations in the
rescaled velocity and temperature probably result from the numerical viscosity, which is not negligible
around the sharp gradient near the second junction.
On the other hand, if we keep all the parameters as in Section 4.1 but
n
=
n0
with
n0
=
0.1 m
2
/Vs, and is again a parameter. Correspondingly, we have =
0
and a xed =
0
, where

0
and
0
are dened above. The limiting system is the drift-diffusion model (18)(20).
The numerical results are displayed in Fig. 13. The temperature differs less from the lattice
temperature, when decreasing the mobility constant. This is clear from the asymptotic limit (see
Section 2.2). At =0.00258, the maximal difference is about 1 K. The rescaled velocity converges to a
prole. As a result, the total energy density w converges to constant, as the main contribution comes from
the thermal energy. The energy density therefore converges to a prole similar to that of the concentration.
We solve the drift-diffusion model (18)(20) numerically employing a relaxation scheme, and observe
a nice agreement with the numerical limit of the hydrodynamic model (see Fig. 14) except some small
oscillations around the junctions, due to numerical viscosity.
A. Jngel, S. Tang / Applied Numerical Mathematics 43 (2002) 229252 247
Fig. 8. Numerical solutions for different lattice temperatures. Dotted: T
0
=300 K, dash-dot: T
0
=200 K, dashed: T
0
=100 K,
solid: T
0
=80 K, heavy-dotted: T
0
=60 K.
Fig. 9. Numerical solutions for different type of boundary conditions. Dotted: Dirichlet boundary conditions, solid: Neumann
boundary conditions.
248 A. Jngel, S. Tang / Applied Numerical Mathematics 43 (2002) 229252
Fig. 10. Numerical solutions for different type of boundary conditions (50 nm-channel). Dotted: Dirichlet boundary conditions,
solid: Neumann boundary conditions.
Fig. 11. Approximation to the energy-transport model. Dotted: (, ) =(0.10315, 2.0116), dash-dot: (, ) =(0.05158, 1.0058),
dashed: (, ) =(0.02063, 0.4023), solid: (, ) =(0.01032, 0.2012), heavy-dotted: (, ) =(0.00516, 0.1006).
A. Jngel, S. Tang / Applied Numerical Mathematics 43 (2002) 229252 249
Fig. 12. Comparison between the stationary energy-transport solution and the hydrodynamic model with (, ) =
(0.00516, 0.1006). Solid: energy-transport model solution, heavy dotted: hydrodynamic model solution. (a) Electron density,
(b) rescaled velocity, (c) temperature, (d) electric eld.
5. Conclusions
In this paper, we have applied a relaxation scheme to simulate the hydrodynamic model for semi-
conductor devices. We have demonstrated its accuracy and efciency through numerical experiments.
With this scheme, we have further investigated interesting features of the system when varying different
parameters and the geometry. The dependence of the velocity overshoot on the low eld mobility and
the channel length has been studied. It turns out that for sufciently small mobilities, the hump near
the rst junction disappears. Numerical limits, as well as a theoretical study by formal expansion, have
been performed, yielding the energy-transport model and the drift-diffusion model, for different limit
processes.
The relaxation approach has recently been applied in higher space dimension, and will be presented
in a forthcoming paper [26].
250 A. Jngel, S. Tang / Applied Numerical Mathematics 43 (2002) 229252
Fig. 13. Approximation to the drift-diffusion model with = 2.0116. Dotted: = 0.10315, dash-dot: = 0.05158, dashed:
=0.01032, solid: =0.00516, heavy-dotted: =0.00258.
Acknowledgements
The authors acknowledge partial support from the Gerhard-Hess Program of the Deutsche Forschungs-
gemeinschaft, grant JU 359/3-1, and from the AFF Project of the University of Konstanz, grant 4/00. The
rst author has been supported partially by the TMR Project Asymptotic Methods in Kinetic Theory,
grant ERB-FRMXCT-970157. The second author has been partially supported by Chinese Special Funds
for Major State Basic Research Project, NSFC under grant 10002002, and a DAAD grant.
A. Jngel, S. Tang / Applied Numerical Mathematics 43 (2002) 229252 251
Fig. 14. Comparison between the drift-diffusion solution and the hydrodynamic model with (, ) =(0.00258, 2.0116). Solid:
drift-diffusion model solution, heavy dotted: hydrodynamic model solution. (a) Electron density, (b) rescaled velocity.
We would like to thank Professor Jos A. Carrillo for interesting discussions, as well as the help of
Professor Irene M. Gamba. We also thank Mr. Stefan Holst for providing the stationary energy-transport
model solution for comparison.
References
[1] G. Ali, D. Bini, S. Rionero, Global existence and relaxation limit for smooth solutions to the EulerPoisson model for
semiconductors, SIAM J. Appl. Math. 32 (2000) 572587.
[2] M. Anile, J.A. Carrillo, I.M. Gamba, C.-W. Shu, Approximation of the BTE by a relaxation-time operator: Simulations for
a 50 nm-channel Si Diode, VLSI Des. 13 (2001) 349354.
[3] M. Anile, V. Romano, G. Russo, Extended hydrodynamic model of carrier transport in semiconductors, SIAM J. Appl.
Math. 61 (2000) 74101.
[4] D. Aregba-Driollet, R. Natalini, Discrete kinetic schemes for multidimensional system of conservation laws, SIAM J.
Numer. Anal. 37 (2000) 19732004.
[5] D. Aregba-Driollet, R. Natalini, S. Tang, Numerical study of diffusive BGK approximations for nonlinear systems of
degenerate parabolic equations, Preprint, 2000.
[6] G. Baccarani, M. Wordeman, An investigation on steady-state velocity overshoot in silicon, Solid-State Electr. 29 (1982)
970977.
[7] N. Ben Abdallah, P. Degond, On a hierarchy of macroscopic models for semiconductors, J. Math. Phys. 37 (1996) 3308
3333.
[8] N. Ben Abdallah, P. Degond, S. Gnieys, An energy-transport model for semiconductors derived from the Boltzmann
equation, J. Statist. Phys. 84 (1996) 205231.
[9] K. Bltekjr, Transport equations for electrons in two-valley semiconductors, IEEE Trans. Electr. Dev. 17 (1970) 3847.
[10] R. Caisch, S. Jin, G. Russo, Uniformly accurate schemes for hyperbolic systems with relaxation, SIAMJ. Numer. Anal. 34
(1997) 246281.
[11] C. Cercignani, I. Gamba, J. Jerome, C.-W. Shu, Device Benchmark comparisons via kinetic, hydrodynamic, and high-eld
models, Comput. Methods Appl. Mech. Engrg. 181 (2000) 381392.
[12] G.-Q. Chen, J. Jerome, B. Zhang, Existence and the singular relaxation limit for the inviscid hydrodynamic energy model,
in: J. Jerome (Ed.), Modelling and Computation for Application in Mathematics, Science, and Engineering, Clarendon
Press, Oxford, 1998.
252 A. Jngel, S. Tang / Applied Numerical Mathematics 43 (2002) 229252
[13] Z. Chen, B. Cockburn, J. Jerome, C.-W. Shu, Mixed-RKDG nite element methods for the 2-D hydrodynamic model for
semiconductor device simulation, VLSI Des. 3 (1995) 145158.
[14] P. Degond, A. Jngel, P. Pietra, Numerical discretization of energy-transport models for semiconductors with non-parabolic
band structure, SIAM J. Sci. Comput. 22 (2000) 9861007.
[15] M. Fischetti, S. Laux, Monte Carlo study of electron transport in silicon inversion layers, Phys. Rev. B 48 (1993) 2244
2274.
[16] C. Gardner, Numerical simulation of a steady-state electron shock wave in a submicron semiconductor device, IEEE Trans.
Electr. Dev. 38 (1991) 392398.
[17] I. Gasser, R. Natalini, The energy-transport and the drift-diffusion equations as relaxation limits of the hydrodynamic
model for semiconductors, Quart. Appl. Math. 57 (1999) 269282.
[18] S. Holst, A. Jngel, P. Pietra, A mixed nite-element discretization of the energy-transport equations for semiconductors,
SIAM J. Sci. Comp., in press.
[19] J. Jerome, C.-W. Shu, Energy models for one-carrier transport in semiconductor devices, in: W. Coughran, J. Colde,
P. Lloyd, J. White (Eds.), Semiconductors, Part II, in: IMA Vol. in Math. Appl., Vol. 59, Springer, New York, 1994,
pp. 185207.
[20] X. Jiang, A streamline-upwinding/PetrovGalerkin method for the hydrodynamic semiconductor device model, Math.
Models Meth. Appl. Sci. 5 (1995) 659681.
[21] S. Jin, J. Liu, The effects of numerical viscosities I: slowly moving shocks, J. Comput. Phys. 126 (1996) 373389.
[22] S. Jin, Z.P. Xin, The relaxation schemes for systems of conservation laws in arbitrary space dimensions, Comm. Pure Appl.
Math. 48 (1995) 235278.
[23] S. Junca, M. Rascle, Relaxation of the isothermal EulerPoisson system to the drift-diffusion equations, Quart. Appl.
Math. 58 (2000) 511521.
[24] A. Jngel, Asymptotic limits in macroscopic plasma models, in: Proceedings of the IMA Workshop, Minneapolis, 2001,
to appear.
[25] A. Jngel, Quasi-Hydrodynamic Semiconductor Equations, in: Progress in Nonlinear Differential Equations, Birkhuser,
Basel, 2001.
[26] A. Jngel, S. Tang, Article in preparation, 2001.
[27] R. Levesque, Numerical Methods for Conservation Laws, Birkhuser, Basel, 1990.
[28] T.P. Liu, Hyperbolic conservation laws with relaxation, Comm. Math. Phys. 108 (1987) 153175.
[29] P.A. Markowich, C.A. Ringhofer, C. Schmeiser, Semiconductor Equations, Springer, Berlin, 1990.
[30] R. Natalini, Recent mathematical results on hyperbolic relaxation problems, in: H. Freisthler (Ed.), Analysis of Systems
of Conservation Laws, Chapman and Hall, London, 1999, pp. 128198.
[31] R. Natalini, S. Tang, Discrete kinetic models for dynamic phase transitions, Comm. Appl. Nonlinear Anal. 7 (2000) 132.
[32] V. Romano, G. Russo, Numerical solution for hydrodynamic models of semiconductors, Preprint, Universit dellAquila,
Italy, 1997.
[33] W. van Roosbroeck, Theory of ow of electrons and holes in germanium and other semiconductors, Bell Syst. Techn. J. 29
(1950) 560607.
[34] M. Rudan, F. Odeh, Multi-dimensional discretization scheme for the hydrodynamic model of semiconductor devices,
COMPEL 5 (1986) 149183.
[35] D. Scharfetter, H. Gummel, Large signal analysis of a Silicon Read diode oscillator, IEEE Trans. Electr. Dev. ED-16 (1969)
6477.
[36] S. Selberherr, Analysis and Simulation of Semiconductor Devices, Springer, Berlin, 1984.
[37] C.-W. Shu, Essentially non-oscillatory and weighted essentially non-oscillatory schemes for hyperbolic conservation laws,
ICASE Report No. 97-65, NASA Langley Research Center, Hampton, VA, 1997.
[38] C.-W. Shu, S. Osher, Efcient implementation of essentially non-oscillatory shock capturing schemes II, J. Comput.
Phys. 83 (1989) 3278.
[39] R. Stratton, Diffusion of hot and cold electrons in semiconductor barriers, Phys. Rev. 126 (1962) 20022014.

Vous aimerez peut-être aussi