You are on page 1of 161

Transport Equations for Semiconductors

Prof. Dr. Ansgar J ungel


Institut f ur Mathematik
Johannes Gutenberg-Universitat Mainz
Winter 2004
Preliminary Lectures Notes (Feb. 2005)
Contents
1 Introduction 3
2 Basic Semiconductor Physics 8
2.1 Semiconductor crystals . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
2.2 The semi-classical picture . . . . . . . . . . . . . . . . . . . . . . . . . . . 21
2.3 The k p method . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28
2.4 Semiconductor statistics . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33
3 Classical Kinetic Models 46
3.1 The Liouville equation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 46
3.2 The Vlasov equation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 52
3.3 The Boltzmann equation . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58
3.4 The bipolar Boltzmann model . . . . . . . . . . . . . . . . . . . . . . . . . 68
4 Classical Fluid Models 72
4.1 Derivation of the drift-diusion equations . . . . . . . . . . . . . . . . . . . 72
4.2 Derivation of the hydrodynamic equations . . . . . . . . . . . . . . . . . . 84
4.3 Derivation of the spherical harmonics expansion equations . . . . . . . . . 91
4.4 Derivation of the energy-transport equations . . . . . . . . . . . . . . . . . 103
4.5 Relaxation-time limits . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 114
5 Quantum Kinetic Models 119
5.1 The Schrodinger equation . . . . . . . . . . . . . . . . . . . . . . . . . . . 119
1
5.2 The quantum Liouville equation . . . . . . . . . . . . . . . . . . . . . . . . 121
5.3 The quantum Vlasov and Boltzmann equation . . . . . . . . . . . . . . . . 125
5.4 Open quantum systems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 131
6 Quantum Fluid Models 136
6.1 Zero-temperature quantum hydrodynamic equations . . . . . . . . . . . . . 136
6.2 Quantum hydrodynamics and the Schrodinger equation . . . . . . . . . . . 138
6.3 Quantum hydrodynamics and the Wigner equation . . . . . . . . . . . . . 141
6.4 Derivation of the quantum drift-diusion equations . . . . . . . . . . . . . 150
2
1 Introduction
The modern computer and telecommunication industry relies heavily on the use and devel-
opment of semiconductor devices. The rst semiconductor device (a germanium transistor)
has been built in 1947 by Bardeen, Brattain and Shockley, who have been awarded the
Nobel prize in 1956. In the following decades, a lot of dierent devices for special appli-
cations have been invented; for instance, semiconductor lasers, solar cells, light-emitting
diodes (LED), metal-oxide semiconductor (MOS) transistors, quantum dots, to name only
a few.
A very important fact of the success of the semiconductor technology is that the device
length is much smaller than that of previous electronic devices (like tube transistors).
The rst transistor of Bardeen, Brattain and Shockley had a characteristic length (the
emitter-collector length) of 20 m, compared to the size of a few centimeters of a tube
transistor. The rst Intel processor 4004, built in 1971, consisted of 2250 transistors, each
of them with a characteristic length of 10 m. This length has been reduced to 90 nm
for the transistors in the Pentium 4 processor (put on the market in June 2004). Modern
quantum-based devices (like tunneling diodes) have structures of only a few nanometer
length. Clearly, on such scales, the physical phenomena have to be described by equations
from quantum mechanics.
Usually, a semiconductor device can be considered as a device which needs an input (an
electronic signal or light) and produces an output (light or an electronic signal). The device
is connected to the electric circuit by contacts at which a voltage (potential dierence)
is applied. We are mainly interested in devices which produce an electronic signal, for
instance the current of electrons through the device, generated by the applied bias. In
this situation, the input parameter is the applied voltage and the output parameter is the
electron current through one contact. The relation between these two physical quantities
is called current-voltage characterestic. It is a curve in the two-dimensional current-voltage
space and does neither need to be a monotone mapping nor a function (but a relation).
The main objective of these lecture notes is to derive mathematical models which
describe the electron ow through a semiconductor device due to the application of a
voltage. Depending on the device structure, the main transport phenomena may be very
dierent, caused by diusion, drift, scattering, or quantum mechanical eects. Moreover,
usually a large number of electrons is owing through a device such that a particle-like
description using uid-dynamical or kinetic equations seems to be appropriate. On the
other hand, electrons in a semiconductor crystal are quantum mechanical objects such
that a wave-like description using the Schrodinger equation is necessary. For this reason,
we have to devise dierent mathematical models which are able to describe the important
physical phenomena for a particular situation or for a particular device. Moreover, since in
3
some cases we are not interested in all the available physical information, we need simpler
models which help to reduce the computation cost in the numerical simulations.
This leads to a hierarchy of semiconductor models. Roughly speaking, we distin-
guish three classes of semiconductor models: quantum models, kinetic models, and uid-
dynamical (macroscopic) models. In order to give some avor of these models, we explain
these three view points, quantum, kinetic, and uid-dynamical, in the following in a sim-
plied framework.
The uiddynamical view. Consider an ensemble of electrons in a spatial domain
R
3
under the inuence of an electric eld. We wish to nd equations which describe the
evolution of the number N

(t) of electrons and the electric eld. For this, we rst introduce
the electron density n(x, t) at time t by
N

(t) =
_

n(x, t) dx.
Another important physical quantity is the electron current density J(x, t). We assume
that it is given as the sum of the drift current, q
k
nE, and the diusion current, qD
n
n,
J = qD
n
n + q
n
nE, (1.1)
where q is the elementary charge,
n
the mobility and D
n
the diusivity of the electrons.
Physically, it is reasonable to assume that the temporal change of the electron number is
equal to the (normal component of the) current ow through the semiconductor boundary
dN

dt
=
1
q
_

J ds,
where and is the exterior unit normal to . The divergence theorem implies
_

n
t
dx =
dN

dt
=
1
q
_

divJ dx for all .


Therefore,
n
t

1
q
divJ = 0 in . (1.2)
This equation expresses the conservation of mass.
It remains to nd an equation for the electrostatic potential. We start with the Maxwell
equations
curl E = 0, divD = in R
3
, (1.3)
valid for vanishing magnetic elds. Here, D is the displacement vector and the total space
charge. The rst equation provides the existence of a potential V such that E = V .
4
The function V is called the electrostatic potential. We assume that the displacement
vector D and the electric eld E are related by
D =
s
E,
where
s
is the semiconductor permittivity. Generally,
s
is a matrix and depends on
the spatial variable. In an isotropic homogeneous semiconductor,
s
is a scalar. The total
space charge is given by
= qn + qC(x),
where C(x) is the concentration of the xed charged background ions in the semiconduc-
tor crystal and will be subject of later investigation (see Section 2.4). Thus, the second
equation in (1.3) gives

s
V = div(
s
V ) = divD = q(n C(x)), (1.4)
if
s
is a constant scalar (the semiconductor permittivity). This equation is called the
Poisson equation.
We have motivated that the electron density n(x, t) and the electrostatic potential
V (x, t) are solutions of (1.1), (1.2), and (1.4). This set of equations,
n
t
div(D
n
n
n
nV ) = 0,
s
V = q(n C(x)), x R
3
, t > 0,
is called the drift-diusion model and is one of the most important semiconductor models.
The kinetic view. Consider a single electron in the semiconductor and interpret it as
a particle at position x(t) with velocity v(t). These functions are a solution of Newtons
equations
dx
dt
= v, m
dv
dt
= F, t > 0, (1.5)
x(0) = x
0
, v(0) = v
0
,
where m is the electron mass and F a force given by F = qE. The state of the electron
is given by a distribution function f(x, v, t), i.e. a probability density in the (x, v)-phase
space. More precisely,
_
B
f(x, v, t) dxdv
is the probability at time t to nd the electron in the phase space set B. It is reasonable
to assume that f(x, v, t) does not change along the trajectory (x(t), v(t)) of the electron,
f(x(t), v(t), t) = f(x
0
, v
0
, 0) for all t > 0.
5
This implies that
0 =
d
dt
f(x(t), v(t), t) =
f
t
+
x
f
dx
dt
+
v
f
dv
dt
,
and, by (1.5),
f
t
+ v
x
f
q
m
E
v
f = 0 in R
3
. (1.6)
This equation is called the Liouville equation.
Macroscopic quantities, like the electron density n(x, t) and the electron current den-
sity, are dened in terms of the distribution function by
n(x, t) =
_
R
3
f(x, v, t) dv,
J(x, t) = q
_
R
3
f(x, v, t)v dv.
The macroscopic particle velocity u(x, t) is then given by
u(x, t) =
_
f(x, v, t)v dv
_
f(x, v, t) dv
=
J(x, t)
qn(x, t)
. (1.7)
We notice that equation (1.2) can be easily derived from the Liouville equation. Indeed,
integrating (1.6) over v R
3
yields, by the divergence theorem,
0 =
_
R
3
_
f
t
+ div
x
(vf)
q
m
div
v
(Ef)
_
dv
=

t
_
R
3
f dv + div
x
_
R
3
vf dv
=
n
t

1
q
div
x
J.
The quantum view. We consider now a single electron interpreted as a wave. Then
it is described by the complex-valued wave function (x, t) which is a solution of the
Schrodinger equation
i

t
=

2
2m
qV (x), t > 0, (x, 0) =
0
(x), x R
3
. (1.8)
Here, i is the complex unit with i
2
= 1, = h/2 the reduced Planck constant, and V (x)
the (time-independent) electrostatic potential. The measurable, macroscopic quantities
n(x, t) and J(x, t) are dened by
n(x, t) = [(x, t)[
2
, J(x, t) =
q
m
Im(), (1.9)
6
where Im(z) denotes the imaginary part of a complex number z and z its conjugate
number.
Usually, the Schrodinger equation (1.8) is solved by the ansatz
(x, t) = w(t)u(x).
Inserting this expression into (1.8) and dividing by gives
i
w
t
w
=

2
2m
u
u
qV (x).
The left-hand side only depends on t, the right-hand side only on x. Therefore, both sides
must be constant. We call this constant E (since physically, it has the unit of an energy).
The solution of
i
dw
dt
= Ew
is given by w(t) = exp(iEt/) (neglecting the integration constant which is put into
u(x)). The function u(x) solves

2
2m
u qV (x)u = Eu in R
3
. (1.10)
This is the stationary Schrodinger equation for a single electron with energy E. Mathe-
matically, (1.10) is an eigenvalue problem and solutions to (1.8) are eigenfunctions with
eigenvalues E. Then, the solution to (1.8) is given by
(x, t) = e
iEt/
u(x),
which oscillates in time with a frequency = E/.
The above models are all stated or motivated in a very simplied situation. For a more
precise description, the following questions have to be answered:
How can the above models be derived (and not only motivated) from basic princi-
ples?
How can the motion of many electrons be modeled?
What is the inuence of the semiconductor crystal on the motion of the electrons?
How can collisions of the electrons with the crystal atoms or with other particles
can be taken into account?
Do there exist relations between the above models?
These questions will be answered (at least partially) in the following chapters.
7
2 Basic Semiconductor Physics
In this chapter we present a short summary of the physics and main properties of semi-
conductors. Only those subjects relevant to the subsequent chapters are included here. We
refer to [1, 3, 40, 61] for introductory textbooks of solid-state and semiconductor physics
and to [11, 37, 45, 50, 65] for more advanced expositions.
2.1 Semiconductor crystals
What is a semiconductor? Historically, the term semiconductor has been used to de-
note solid materials whose conductivity is much larger than that for insulators but much
smaller than that for metals, measured at room temperature. A modern and more pre-
cise denition is that a semiconductor is a solid with an energy gap larger than zero and
smaller than about 4eV (electron volt). Metals have no energy gap, whereas it is usually
larger than 4eV in insulators. In order to understand the notion energy gap, we have
to introduce to the crystal structure of solids.
A solid is made of an innite three-dimensional array of atoms arranged according to
a lattice
L = n
1
a
1
+ n
2
a
2
+ n
3
a
3
: n
1
, n
2
, n
3
Z R
3
,
where a
1
, a
2
, a
3
are the basis vectors of L, called primitive vectors of the lattice (see
Figure 2.1). The set L is also called a Bravais lattice. The lattice atoms generate a periodic
electrostatic potential V
L
,
V
L
(x + y) = V
L
(x) for all x R
3
, y L,
which is the superposition of the Coulomb potentials
V
j
(x) =
1
4
q
2
[x x
j
[
of the crystal atoms located at x = x
j
(see Figure 2.2). The state of an electron moving in
this periodic potential is described by an eigenfunction (x) of the stationary Schrodinger
equation (1.10):

2
2m
qV
L
(x) = E in R
3
, (2.1)
where : R
3
C is the (stationary) wave function, and as in Chapter 1, = h/2 is
the reduced Planck constant, m the electron mass (at rest), q the elementary charge, and
E the energy (or the eigenvalue corresponding to ). We illustrate this equation and its
solutions by two examples.
8
a
1
a
2
Figure 2.1: Illustration of a two-dimensional lattice L.
0
x
V
j1
(x)
V
L
(x)
V
j+1
(x)
x
j1
x
j
x
j+1
x
j+2
atom atom atom atom
Figure 2.2: Potentials V
j
(x) of a single atom at x = x
j
and net potential V
L
(x) of a
one-dimensional crystal lattice.
Example 2.1 (Motion of a free electron)
Consider a free electron moving in a one-dimensional vacuum, i.e. V
L
(x) = 0 for all x R.
Then the eigenfunctions of

2
2m

tt
= E in R (2.2)
are given by

k
(x) = Ae
ikx
+ Be
ikx
, x R,
where k
2
= 2mE/
2
, and the eigenvalues are
E = E(k) =

2
k
2
2m
, k R.
To be precise, k is complex but the purely imaginary part i ( R) leads to solutions
of the type exp(ikx) = exp(x) for x R. The integral
_
R
[(x)[
2
dx
9
is the particle mass (or particle number) and should be nite. However,
_
R
[exp(x)[
2
dx
is never nite. Therefore, only k R gives physically reasonable solutions.
We have found that the eigenvalue problem (2.2) has innitely many solutions parametrized
by k R and corresponding to dierent energies E(k). The functions exp(ikx) are called
plane waves. Thus, the eigenstates of a free particle are plane waves.
Example 2.2 (Innite square-well potential)
The innite square-well potential is a one-dimensional structure of length L with a van-
ishing potential inside the well and with an innite potential at its boundaries. As the
potential is conning an electron to the inner region, we have to solve the Schrodinger
equation (2.1) in the interval (0, L) with boundary conditions
(0) = (L) = 0
and potential V
L
(x) = 0 for x (0, L) (see Figure 2.3). The eigenfunctions to (2.1) are as
in Example 2.1
(x) = Ae
ikx
+ Be
ikx
with k
2
= 2mE/
2
and k R. The constants A and B are determined by the boundary
conditions:
0 = (0) = A + B, 0 = (L) = Ae
ikL
+Be
ikL
.
Inserting the rst equation in the second one gives
0 = A(e
ikL
e
ikL
) =
A
2i
sin kL.
Disregarding the trivial solution A = 0 (and hence 0), the parameter k should be
such that sin kL = 0, i.e., kL is a multiple of . Thus
k =
n
L
, n N
0
.
The eigenfunctions become

k
(x) =

Asin kx, x [0, L],
where

A = A/2i, with eigenvalues
E(k) =

2
k
2
2m
.
10

x
0 L
`

Figure 2.3: Innite square-well potential.


The integration constant

A R can be determined by assuming that
_
L
0
[
k
(x)[
2
dx = 1
holds. A simple computation shows that

A =

2L. The system only allows discrete energy


states. In particular, the parameter k can only take discrete values.
Since the lattice potential V
L
is periodic, one might hope that the whole-space Schro-
dinger problem (2.1) can be reduced to an eigenvalue problem on a cell of the lattice.
Blochs theorem states that this is indeed possible. Before we can formulare the result,
we need some denitions.
Denition 2.3 [49, p. 12f.]
(1) The reciprocal lattice (or dual lattice) L

of L is dened by
L

= n
1
a

1
+ n
2
a

2
+n
3
a

3
: n
1
, n
2
, n
3
Z,
where the primitive vectors a

1
, a

2
, a

3
R
3
are uniquely given by the relation
a
m
a

n
= 2
mn
, m, n = 1, 2, 3. (2.3)
(2) The connected set D R
3
is called primitive cell of L (or L

) if the volume of D
equals the volume of the parallelepiped spanned by the basis vectors of L (or L

),
vol D =a
1
(a
2
a
3
) (or vol D =a

1
(a

2
a

3
)),
and if the whole space R
3
is covered by the union of translates of D by the primitive
vectors (see Figure 2.4).
(3) The (rst) Brillouin zone B R
3
is the primitive cell of the reciprocal lattice L

which consists of all points being closer to the origin than to any other point of L

(see Figure 2.4):


B = k R
3
: [k[ min

[k + [, L

, ,= 0.
11

a
1
-
0

1
a
2

2
B
q
6
Figure 2.4: The primitive vectors of a two-dimensional lattice L and its reciprocal lattice
L

and the Brillouin zone B.


We give some explanations of the above denition. What is the meaning of the reci-
procal lattice? The reciprocal lattice vectors and the direct lattice vectors can be seen as
conjugate variables, like time and frequency are conjugate variables in signal analysis. In
fact, let x L and k L

be given such that


x =
3

m=1

m
a
m
and k =
3

n=1

n
a

n
,
where
m
,
n
Z. Then, by (2.3),
e
ikx
= exp
_
i
3

m,n=1

n
2
mn
_
= exp
_
2i
3

m=1

n
_
= 1. (2.4)
As x has the dimension of length, k has the dimension of inverse length and therefore,
k is called a wave vector. (More precisely, k is called a pseudo-wave vector; see below).
Physically, the reciprocal lattice appears in X-ray diraction experiments on crystals. It
can be shown that the peaks of intensity of the reected X-ray are obtained when the
change in momentum k of the X-ray wave is an element of the reciprocal lattice [11, p.
404]. This allows to determine the structure of the crystal lattice.
How can the Brillouin zone of the reciprocal latice be constructed? Mathematically,
the primitive vectors a

n
of the Brillouin zone are given as the inverse of the matrix
A whose columns are the vectors a
m
. More precisely, let a
n
= (a
1n
, a
2n
, a
3n
)

, a

m
=
(a

1m
, a

2m
, a

3m
)

R
3
and A

= (a

1
, a

2
, a

3
), A = (a
1
, a
2
, a
3
) R
33
. Then (2.3) implies
12
that
(A

)
mn
=
3

j=1
a
jm
a

jn
=a
m
a
n
= 2
mn
and thus A

= 2I, where I is the unit matrix of R


33
. Hence
A

= 2(A

)
1
= 2(A
1
)

. (2.5)
Graphically, the Brillouin zone can be constructed as follows. Draw arrows from a lattice
point to its nearest neighbors and cut them in half. Then the planes through these mid
points are perpendicular to the arrows from the surface of the (bounded) Brillouin zone.
In two space dimensions, the Brillouin zone is a hexagon or a square (see Figure 2.4). In
three space dimensions, the zone is a polyhedron (a capped octahedron).
Lemma 2.4 The volumes of a primitive cell D and its Brillouin zone B are related to
vol B =
(2)
3
vol D
.
Proof: With the above notations,
vol D =a
1
(a
2
a
3
) = det(a
1
, a
2
, a
3
) = det A,
vol B = det(a

1
, a

2
, a

3
) = det A

and hence, by (2.5),


vol B = det[2(A
1
)

] = (2)
3
det(A
1
) =
(2)
3
det A
=
(2)
3
vol D
.
Now we can formulate the Bloch theorem.
Theorem 2.5 (Bloch)
Let V
L
be a periodic potential, i.e., V
L
(x + y) = V
L
(x) for all x R
3
and y L (the
Bravais lattice). Then the eigenfunctions of

2
2m
qV
L
(x) = E in R
3
(2.6)
can be written as
(x) = e
ikx
u(x) (2.7)
for some k B (the Brillouin zone) and some function u(x) satisfying u(x) = u(x + y)
for all x R
3
, y L.
This theorem asserts that any eigenfunction of the Schrodinger equation is the product
of a plane wave e
ikx
and a function having the periodicity of the lattice L. Before we
can give a formal proof of this result, we need some preparations. In the following we
do neither specify the underlying (Hilbert) space nor the domain of denition of the
considered operators in order to simplify the presentation.
13
Denition 2.6 Let T be an operator, dened on some Hilbert space X with scalar product
(, ).
(1) The adjoint operator T

is formally dened by
(Tx, y) = (x, T

y) for all x, y X.
(2) The operator T is called normal if T

T = TT

.
Example 2.7 (1) Dene the translation operator T
a
for some a R
3
by
(T
a
)(x) = (x +a), x R
3
, (2.8)
for L
2
(R
3
; C), where the scalar product is given by
(, ) =
_
R
3
(x)(x) dx.
We claim that T
a
is normal. From
(, T
a
) =
_
R
3
(x)(x +a) dx =
_
R
3
(y a)(y) dy = (T
a
, )
follows that T

a
= T
a
. Furthermore,
T
a
T
a
= = T
a
T
a
,
and thus, (T
a
)
1
= T
a
= T

a
. This implies that T

a
T
a
= I = T
a
T

a
, where I is the
identity operator. Hence T
a
is normal. (In fact, we have even shown that T
a
is unitary,
i.e. T

a
= (T
a
)
1
.)
Next we claim that the eigenvalues of T
a
are given by = e
i
for R. To see this,
let T
a
= and set ||
2
= (, ). Then
[[
2
||
2
= ||
2
= |T
a
|
2
=
_
R
3
[(x + a)[
2
dx =
_
R
3
[(x)[
2
dx = ||
2
,
and thus, [[ = 1 or = e
i
with R.
(2) We state without proof that the Hamilton operator H, dened by
H =

2
2m
qV
L
(x) (2.9)
for appropriate , is also normal. (In fact, H is self-adjoint, i.e. H

= H.)
Theorem 2.8 (Spectral theorem for normal operators)
Let T be a normal operator on the Hilbert space X. Then there exists a orthonormal basis
of eigenvectors of T.
14
Theorem 2.9 Let S and T be two normal operators on the Hilbert space X such that
ST = TS. Then there exists an (orthonormal) basis of X whose elements are eigenvectors
for both S and T.
In Example 2.7 and Theorems 2.8 and 2.9 we have simplied the situation. To be
precise, the Hamilton operator is dened for all in a dense subset of the underlying
Hilbert space and the operators in Theorems 2.8 and 2.9 have to be compact (i.e., if the
sequence (x
n
) is bounded then (Tx
n
) and (Sx
n
) contain convergent subsequences).
We are now able to turn to the (formal) proof of Blochs theorem.
Proof of Theorem 2.5: By Example 2.7, the translation operator (2.8) and the Hamilton
operator (2.9) are normal. Moreover, since for all a L,
(T
a
H)(x) = T
a
_

2
2m
qV
L

_
(x) =

2
2m
(x + a) qV
L
(x + a)(x + a)
=

2
2m
(x + a) qV
L
(x)(x + a) = H((x + a))
= (HT
a
)(x),
by Theorem 2.9, there exists a basis of eigenfunctions for both T
a
and H. By Exam-
ple 2.7 (1), for given primitive vector a
j
L, there exists
j
R such that
T
a
j
= e
i
j
, (2.10)
and is also an eigenfunction of H. We set
k =
1
2
3

j=1

j
a

j
,
where a

j
L

. Then (2.4) implies that


k a
j
=
j
. (2.11)
We dene u(x) = e
ikx
(x), x R
3
. We have to show that u(x+y) = u(x) for all x R
3
and y L. Since every y L is a linear combination of the a
j
, it is sucient to prove the
periodicity for y =a
j
. We obtain, using (2.10) and (2.11),
u(x) = e
ikx
(x) = e
ikx
(T
a
j
)(x +a
j
) = e
ikx
e
i
j
(x +a
j
)
= e
ikx
e
i
j
e
ik(x+a
j
)
u(x +a
j
) = e
i(
j
+ka
j
)
u(x +a
j
) = u(x +a
j
).
It remains to show that k can be restricted to the Brillouin zone. Let k R
3
and
decompose k = k
B
+, where k
B
B and L

is a point in the reciprocal lattice closest


to k (see Figure 2.5). Then
(x) = e
ikx
u(x) = e
ik
B
x
(x), (2.12)
15
where (x) = e
ix
u(x) satises, in view of (2.4),
(x + y) = e
ix
e
iy
u(x +y) = e
ix
u(x) = (x)
for all x R
3
and y L. Now, the representation (2.12) is of the form as stated in the
theorem.
k

B
k
B
a

1
a

2
Figure 2.5: Illustration of k = k
B
+ with k
B
B and L

for the proof of Theorem 2.5.


Inserting the decomposition (2.7) into the Schrodinger equation (2.6) shows that u(x)
satises the eigenvalue problem

2
2m
(u + 2ik u) +
_

2
2m
[k[
2
qV
L
(x)
_
u = Eu in D, (2.13)
with periodicity conditions
u(x + y) = u(x), x R
3
, y L, (2.14)
since
Eu = e
ikx
E = e
ikx
_

2
2m

_
e
ikx
u
_
qV
L
(x)e
ikx
u
_
=

2
2m
_
u + 2ik u [k[
2
u
_
qV
L
(x)u.
For each k B, there exists a sequence of eigenfunctions u = u
n,k
and eigenvalues
E = E
n
(k) of (2.13)(2.14). (This follows from the self-adjointness of the operator dened
by the left-hand side of (2.13); see [56] for details). In particular, the functions (u
n,k
)
nN
form an orthonormal basis of the underlying Hilbert space (see Theorem 2.8). Then we
can introduce the so-called Bloch functions

n,k
(x) = e
ikx
u
n,k
(x).
16
They satisfy the Schrodinger equation in the primitive cell D of L,

2
2m

n,k
qV
L
(x)
n,k
= E
n
(k)
n,k
in D (2.15)
with pseudo-periodic boundary conditions

n,k
(x + y) = e
iky

n,k
(x), x, x + y D. (2.16)
In some sense, the functions
n,k
are plane waves which are modulated by a periodic
function u
k,n
taking into account the inuence of the atom lattice. This also explains why
k is termed pseudo-wave vector. In fact, k appears in distorted plane waves and is thus
not a real wave vector.
The function k E
n
(k) is called dispersion relation or the n-th enery band. It shows
how the energy of the n-th band depends on the (pseudo-)wave vector k. The union of
ranges of E
n
over n N is not necessarily the whole real line R, i.e., there may exist
energies E

for which there is no n N and no k B such that E


n
(k) = E

. The
connected components of the set of energies with this non-existence property are called
energy gaps. We illustrate this property by the following example.
Example 2.10 (Kronig-Penney model)
The Kronig-Penney model is a simple model representing a one-dimensional single-crystal
lattice (see also [50, Sec. 3.1.2] or [11, Sec. 8.2]). The potential of the lattice atoms is
modeled by the function
V
L
(x) =
_
V
0
if b < x 0
0 if 0 < x a,
and V
L
is extended to R with period a +b:
V
L
(x) = V
L
(x + a +b) for x R,
where a, b > 0, V
0
< 0 (see Figure 2.6). The lattice atoms are supposed to be at the
positions a/2 +n(a +b), n Z.
In order to solve the Schrodinger equation (2.6) we make the Bloch decomposition
(x) = e
ikx
u(x),
where u(x) is a (a +b)-periodic solution of (see (2.13))
u
tt
2iku
t
+k
2
u =
2m

2
(E + qV
L
)u in R. (2.17)
We proceed as in [50, Sec. 3.1.2]. First we solve (2.17) in the interval (0, a). Then V
L
(x) = 0,
x (0, a), and the ansatz u(x) = e
ix
(x R) leads to
(
2
+ 2k + k
2

2
)e
ix
= 0,
17
b 0
V
0
a a +b 2a +b x
Figure 2.6: The periodic square-well potential V
L
(x) of the Kronig-Penney model.
where
=
_
2mE

2
.
The solutions of the above quadratic equation are given by
1/2
= k and therefore,
u
1
(x) = Ae
i(k)x
+Be
i(+k)x
, x (0, a).
In the interval (b, 0) we make again the ansatz u(x) = e
ix
yielding
(
2
+ 2k + k
2

2
)e
ix
= 0,
where
=
_
2m(E qV
0
)

2
.
The solutions are
1/2
= k , and thus
u
2
(x) = Ce
i(k)x
+ De
i(+k)x
, x (b, 0).
We notice that is purely imaginary if E < qV
0
, i.e., the electrons are bound within the
crystal, and is real if E > qV
0
.
The constants A, B, C, D are determined from the interface conditions. We assume
that u is continuously dierentiable and periodic on R:
u
1
(0) = u
2
(0), u
t
1
(0) = u
t
2
(0),
u
1
(a) = u
2
(b), u
t
1
(a) = u
t
2
(b).
We obtain the following four equations:
A + B C D = 0,
( k)A ( + k)B ( k)C + ( + k)D = 0,
Ae
i(k)a
Be
i(+k)a
Ce
i(k)b
De
i(+k)b
= 0,
( k)Ae
i(k)a
( + k)Be
i(+k)a
( k)Ce
i(k)b
+ ( +k)De
i(+k)b
= 0
18
for the unknowns A, B, C, D. This is a homogeneous linear system which has non-trivial
solutions only if the determinant of the coecient matrix vanishes. A very lenghty calcu-
lation shows that this condition is equivalent to

2
+
2
2
sin(a) sin(b) + cos(a) cos(b) = cos(k(a +b)). (2.18)
This equation relates the wave vector k to the energy E through the parameters and .
There are values of E for which there does not exist any k satisfying (2.18). In order
to see this we assume that E < qV
0
such that is purely imaginary and set = i. Since
sin(ix) = i sinh(x) and cos(ix) = cosh(x), (2.18) becomes

2
2
sin(a) sinh(b) + cos(a) cosh(b) = cos(k(a + b)). (2.19)
Using
lim
0
sin(a)

= a,
we obtain for E = 0 (which implies = 0):
a
2
sinh(b) + cosh(b) = cos(k(a + b)).
Since sinh(b) > 0 and cosh(b) > 1, the left-hand side is strictly larger than one and thus,
this equation cannot have a solution. By continuity, there is no solution in a neighborhood
of E = 0.
We can compute the intervals for which no solution exists more easily by simplifying
(2.19). For this, we let the potential barrier width b 0 and the barrier height V
0

such that the product bV
0
R. Then
b =
_
2m(qV
0
E)b
2

2
0, cosh(b) 1
and
(
2

2
)b
2
sinh(b)
b
=
m(qV
0
2E)b

2
sinh(b)
b

mq

2
.
Thus, (2.19) becomes in the limit
f(a) := Q
sin(a)
a
+ cos(a) = cos(ka), (2.20)
where Q = mqa/
2
. Figure 2.7 shows the function f(a). In regions, where [f(a)[ 1,
there exist one or two solutions k of (2.20); in regions with [f(a)[ > 1, no solution k exists.
Every connected subset of [0, )R(E), where R(E) = E
0
0 : k R : E(k) = E
0
,
is an energy gap.
19
10 5 0 5 10
1
0
1
2
3
4
a
f
(


a
)
Figure 2.7: The function f(a) of (2.20) depending on the energy a.

`
`
E
g
E
k
valence band
conduction
band
Figure 2.8: Schematic band structure with energy gap E
g
.
The energy gap separates two energy bands. The nearest energy band below the energy
gap (if it is unique) is called valence band, the nearest energy band above the energy gap
is termed conduction band (see Figure 2.8).
Now we are able to state the denition of a semiconductor: it is solid with an energy
gap whose value is positive and smaller than about 4eV. In Table 2.1 the values of the
energy gaps for some commun semiconductor materials are collected.
Notice that the band structure of real crystals in three space dimensions is much
more complicated than the one-dimensional situation of the Kronig-Penny model. Indeed,
electrons traveling in dierent directions encounter dierent potential patterns, generated
by the lattice atoms, and therefore the E(k) diagram is strictly speaking a function of
the three-dimensional wave vector k. In physics textbooks, usually a projection of the full
E(k) diagram is shown. For instance, Figure 2.9 shows the band structures of silicon and
gallium arsenide. In place of the positive and negative k axes of the one-dimensional case,
20
Material Symbol Energy gap in eV
Silicon Si 1.12
Germanium Ge 0.67
Gallium arsenide GaAl 1.42
Aluminium gallium arsenide Al
0.3
Ga
0.7
As 1.80
Gallium phosphide GaP 2.20
Table 2.1: Energy gaps of selected semiconductors (from [3, Table 28.1] and [45, Fig.
1.14]).
two dierent crystal directions are shown, namely the k = (0, 0, 1)

direction along the


+k axis (also called the line) and the k = (1, 1, 1)

direction along the k axis (also


called the line). The point k = (0, 0, 0) is termed point. The points at the boundary
of the Brillouin zone in the and directions are called L and X points, respectively
(see Figure 2.10).
Conduction
Valence
band
band
E
g
E
g
L
L


X
X
E
n
e
r
g
y
E
n
e
r
g
y
Figure 2.9: Schematic band structure of silicon (left) and gallium arsenide (right) (see [50,
Fig. 3.23] or [65, Fig. 37 and 3.8]).
2.2 The semi-classical picture
The transport of electrons in semiconductors is based on the Schrodinger equation
i

t
=

2
2m
q(V (x) + V
L
(x)),
21
k
1
k
3
L

X
Figure 2.10: Projection of a simplied Brillouin zone on the (k
1
, k
3
) plane and illustration
of the , L, and X points and the and directories.
where V
L
(x) is the (periodic) potential due to the atoms of the crystal lattice and V (x)
is a (non-periodc) potential that is built-in or applied to the semiconductor. The solution
of this equation is extremely dicult such that approximate models need to be used.
One possibility is the semi-classical treatment which describes the carrier dynamics in the
potential V (x) by Newtons laws without explicitly treating the crystal potential V
L
(x).
The inucence of V
L
(x) is indirectly taken into account by the use of the energy band
structure in the description of the velocity and the mass of the carrier ensemble. Therefore,
we will rst motivate two formulas for the mean electron velocity and the eective mass
as functions of the energy bands:
The mean velocity (or group velocity of the wave packet) in the n-th band is given
by
v
n
(k) =
1

k
E
n
(k). (2.21)
The eective mass tensor m

is dened by
(m

)
1
=
1

d
2
E
n
dk
2
. (2.22)
First, we motivate (2.21) by following [11, Sec. 8.1]. We omit the index n in the
following and dene the group velocity by
v(k) =
__
D
[
k
[
2
dx
_
1
_
D
v
k
[
k
[
2
dx, (2.23)
22
where D is a primitive cell of the lattice and v
k
is the particle velocity,
v
k
=
J
k
qn
k
=

m
Im(
k

k
)
[
k
[
2
. (2.24)
The rst equality follows from (1.7) and the second one from (1.9). The wave functions

k
(x) are solutions of (2.15),

2
2m

k
qV
L
(x)
k
= E(k)
k
in D, (2.25)
and can be decomposed by the Bloch theorem (see (2.7)),

k
(x) = e
ikx
u
k
(x). (2.26)
Dierentiating (2.25) with respect to k and using (2.26) gives
(
k
E)
k
=

2
2m

x
(
k
) (qV
L
+ E)
k

k
=

2
2m

x
_
e
ikx

k
u
k
+ ix
k
_
(qV
L
+E)
_
e
ikx

k
u
k
+ ix
k
_
=

2
2m
_

x
(e
ikx

k
u
k
) + 2i
x

k
+ ix
x

k
_
(qV
L
+E)
_
e
ikx

k
u
k
+ ix
k
_
=
_

2
2m

x
(qV
L
+ E)
_
_
e
ikx

k
u
k
_

i
2
m

k
+ix
_

2
2m

x
(qV
L
+ E)
_

k
.
Observing that the last term vanishes in view of (2.25), multiplication of the resulting
equation with
k
and integration over D yields

k
E
_
D
[
k
[
2
dx +
i
2
m
_
D
(
x

k
)
k
dx
=
_
D

k
_

2
2m

x
(qV
L
+ E)
_
_
e
ikx

k
u
k
_
dx
=
_
D
e
ikx

k
u
k
_

2
2m

x
(qV
L
+E)
_

k
dx
= 0,
where we have used integration by parts and again (2.25). The boundary integral in the
integration-by-parts formula vanishes since u
k
is periodic on D. Thus

k
E =
i
2
m
_
D

k

k
dx
_
D
[
k
[
2
dx
.
23
Taking the real part of both sides of this equation gives

k
E =

2
m
Im
_
D

k

k
dx
_
D
[
k
[
2
dx
= v(k),
employing (2.23) and (2.24). This shows (2.21).
The expression (2.21) has some consequences. The change of energy with respect to
time equals the product of a force F and the velocity v
n
:

t
E
n
(k) = Fv
n
(k) =
1
F
k
E
n
(k).
Since, by the chain rule,
t
E
n
(k) =
k
E
n
(k)
t
k, we conclude that
F =
t
(k). (2.27)
Newtons law states that the force equals the time derivative of the momentum p. This
motivates the denition of the crystal momentum
p = k. (2.28)
It should not be confused with the momentum operator p
Q
= i
x
of quantum me-
chanics.
Another consequence of (2.21) is formula (2.22). Indeed, dierentiating (2.21) leads to

t
v
n
=
1

d
2
E
n
dk
2

t
k =
1

2
d
2
E
n
dk
2
F,
using (2.27). The momentum p equals the product of velocity and (eective) mass, p =
m

v
n
. Then, by Newtons law F =
t
p = m

t
v
n
, and we infer that
(m

)
1
=
1

2
d
2
E
n
dk
2
.
This equation is considered as a denition of the eective mass m

. The right-hand side


of this denition is the Hessian matrix, so the symbol (m

)
1
is also a (3 3) matrix.
The eective mass has the advantage that under some conditions, the behavior of the
electrons in a crystal can be described as a free electron gas, for which E(k) =
2
[k[
2
/2m
holds (see Example 2.1). In order to see this, we evaluate the Hessian of E
n
near a local
minimum (of the conduction band), i.e.
k
E
n
(k
0
) = 0. Then d
2
E
n
(k
0
)/dk
2
is a symmetric,
positive matrix which can be diagonalized and the diagonal matrix has positive entries.
We assume that the coordinates are chosen such that d
2
E
n
(k
0
)/dk
2
is already diagonal,
1

2
d
2
E
n
dk
2
(k
0
) =
_
_
_
1/m

1
0 0
0 1/m

2
0
0 0 1/m

3
_
_
_
.
24
Assume that the energy values are shifted in such a way that E
n
(k
0
) = 0. (This is possible
by xing a reference point for the energy.) Let us assume that already E
n
(0) = 0, otherwise
dene

E
n
(k) = E
n
(k + k
0
). If the function k E
n
(k) is smooth, Taylors formula then
implies
E
n
(k) = E
n
(0) +
k
E
n
(0) k +
1
2
k

_
d
2
E
n
dk
2
(0)
_
k +O([k[
3
)
=

2
2
_
k
2
1
m

1
+
k
2
2
m

2
+
k
2
3
m

3
_
+ O([k[
3
),
where k = (k
1
, k
2
, k
3
)

. If the eective masses are equal in all directions, i.e. m

= m

1
=
m

2
= m

3
, we can write, neglecting higher-order terms
E
n
(k) =

2
2m

[k[
2
. (2.29)
This relation is valid for wave vectors k suciently close to a local band minimum (of the
conduction band). The scalar m

is called here the isotropic eective mass. Comparing


this expression with the dispersion relation of a free electron gas,
E(k) =

2
2m
[k[
2
,
we infer that the energy of an electron near a band minimum equals the energy of a free
electron in a vacuum where the (rest) electron mass m is replaced by the eective mass
m

.
The expression (2.29) is referred to as the parabolic band approximation and usually,
the range of wave vectors is extended to the whole space, k R
3
. This simple model
is appropriate for low applied elds for which the carriers are close to the conduction
band minimum. For high applied elds, however, the higher-order terms in the above
Taylor expansion cannot be ignored. In order to account for non-parabolic eects, often
the non-parabolic band approximation in the sense of Kane is used (see [61, Sec. 2.1] or
[50, (1.40)]):
E
n
(1 +E
n
) =

2
2m

[k[
2
, (2.30)
where m

is determined from (2.22) at the conduction band minimum at k = 0,


=
1
E
g
_
1
m

m
_
2
,
and E
g
is the band gap. In Table 2.2 some values for are shown. Formula (2.30) can be
obtained from approximate solutions to (2.13) derived by the so-called k p theory (see
below).
25
Material Si Ge GaAs
in (eV)
1
0.5 0.65 0.64
Table 2.2: Values of the non-parabolicity parameter for some semiconductors (from [45,
Table 1.1].
When we consider the eective mass denition (2.22) near a maximum (of the valence
band), we nd that the Hessian of E
n
is negative denite. This would lead to a negative
eective mass. However, in the derivation of the mean velocity and consequently of the
eective mass, we have used in (2.24) that the charge of the electron is negative. Employing
a positive charge leads again to a positive eective mass. The corresponding particles are
called holes (or defect electrons). Physically, a hole is a vacant orbital in an otherwise
lled valence band. Thus, the current ow in a semiconductor crystal comes from two
sources: the ow of electrons in the conduction band and the ow of holes in the valence
band. It is a convention to consider the motion of the valence band vacancies rather than
the electrons moving from one vacant orbital to the next (see Figure 2.11).
t = 0 :
crystal atom
`
electron

orbital
-

vacancy
t > 0 :
vacancy

electron
Figure 2.11: Motion of a valence band electron to a neighboring vacant orbital or, equiv-
alently, of a hole in the inverse direction.
Close to the bottom k = 0 of the conduction band in an isotropic semiconductor, we
obtain
E
n
(k) = E
c
+

2
2m

e
[k[
2
, (2.31)
whereas near the top k = 0 of the valence band we have
E
n
(k) = E
v


2
2m

h
[k[
2
, (2.32)
where E
c
is the energy at the conduction band minimum, E
v
the energy at the valence
band maximum, m

e
the eective electron mass, and m

h
the eective hole mass. Clearly,
the energy gap E
g
is given by E
g
= E
c
E
v
(see Figure 2.12). Some values for the eective
masses of commun semiconductors can be found in Table 2.3.
26
E
c
E
v
E
g
conduction
band
band
valence
k
Figure 2.12: Schematic conduction and valence bands near the extrema at k = 0.
Material m

e
/m
e
m

h
/m
h
Si 0.98 0.16
Ge 1.64 0.04
GaAs 0.067 0.082
Table 2.3: Relative eective electron and hole masses for some semiconductor material
(from [32, Sec. 2.3.1]). The electron and hole masses at rest are denoted by m
e
and m
h
,
respectively.
Now we come back to the semi-classical picture. In this picture, the motion of an
electron in the n-th band is approximately described by a point particle moving with
velocity v
n
(k). Denoting by (x(t), v
n
(k, t)) the trajectory of an electron in the position-
velocity phase space, we can write, by Newtons law,

t
x = v
n
=
1

k
E
n
,
t
(k) = F,
where F represents a driving force, for instance, F = q
x
V
L
. Notice that band transi-
tions are excluded since the band index n is xed in the equations. In this picture, the
wave packet of an electron is treated as a particle. However, by Heisenbergs uncertainty
principle, the position and the momentum cannot have both sharp values. It is assumed
that the uncertainty in the momentum is so small that the energy of the electron is sharply
dened and that the uncertainty of the position is small compared to the distance over
which the potential varies signicantly.
27
If a non-periodic (external) potential is superimposed to the lattice potential V
L
, the
situation is much more complicated. Indeed, the Schrodinger equation (2.6) cannot be
decomposed into the decoupled Schrodinger equation (2.15) and the energy bands are
now coupled. However, it is usually assumed that the non-periodic potential is so weak
that the coupling of the bands can be neglected, and then the above analysis remains
approximately valid. In particular, we can use the semi-classical equations

t
x = v
n
=
1

k
E
n
,
t
(k) = q
x
V (x, t). (2.33)
This semi-classical treatment will be used in the following chapters.
2.3 The k p method
In the previous section we have seen that the mean velocity and the eective mass of the
electrons in a semiconductor can be computed in the semi-classical picture from the energy
band structure; see formulas (2.21) and (2.22). How can the band structure be computed?
In this section we describe the k p method which allows to derive an approximation of the
energy E
n
(k) close to the bottom of the conduction band or close to the top of the valance
band. The main assumption of this method is that the energy at k = 0 is known. Then
E
n
(k) close to the -point k = 0 can be computed using time-independent perturbation
theory. We proceed in the following as in [11, Sec. 8.7] and [65, Sec. 4.1].
The starting point is the Schrodinger equation (2.13) for the functions u
n,k
of the
Bloch function
n,k
= e
ikx
u
n,k
(see Section 2.1), here written in the form
(H
0
+ H
1
)u
n,k
= E
n
(k)u
n,k
, (2.34)
where
H
0
=

2
2m
qV
L
(x)
is the single-electron Hamiltonian and
H
1
=
i
2
m
k +

2
2m
[k[
2
is considered to be a perturbation of H
0
. Dening the quantum momentum operator
p
Q
= i, we can formally write
H
1
=

m
k p
Q
+

2
2m
[k[
2
,
which explains the name of the k p method.
28
We assume that = [k[ is small compared to one. Notice that for k = 0, the operator
H
0
+ H
1
reduces to the single-electron Hamiltonian. Furthermore, we suppose that the
solutions of the eigenvalue problem
H
0
u
(0)
n
= E
(0)
n
u
(0)
n
in D, (2.35)
where D is the primitive cell, together with periodic boundary conditions are known. Since
the operator H
0
is real, also the eigenfunctions u
(0)
n
are real. We will show the following
result.
Theorem 2.11 Let the solutions (u
(0)
n
) to (2.35) form a non-degenerate orthonormal
basis of L
2
(D, C) (i.e., all eigenspaces are one-dimensional). Then, up to second order in
,
E
n
(k) = E
(0)
n
+

2
2
k

(m

)
1
k (k 0), (2.36)
where the matrix (m

)
1
consists of the elements 1/m

j
with
m
m

j
=
j

2
2
m

q,=n
P
qnj
P
nq
E
(0)
q
E
(0)
n
(2.37)
and
P
qnj
=
_
u
(0)
q
,
u
(0)
n
x
j
_
=
_
D
u
(0)
q
u
(0)
n
x
j
dx.
Notice that the one-dimensionality of the eigenspaces implies that E
(0)
q
,= E
(0)
n
for all
q ,= n and so, (2.37) is dened. The symbol (, ) denotes the scalar product on L
2
(D; C).
Proof: We apply a perturbation method to (2.34) (see [11, Sec. 4.1] or [65, app. C.1.1]).
For this, we develop
u
n,k
= u
(0)
n
+ u
(1)
n
+
2
u
(2)
n
+ , E
n
(k) = E
(0)
n
+ E
(1)
n
+
2
E
(2)
n
+ .
Inserting these expressions into (2.34) and equating terms with the same order of leads
to

0
: H
0
u
(0)
n
= E
(0)
n
u
(0)
n
, (2.38)

1
: H
0
u
(1)
n
+ H
1
u
(0)
n
= E
(0)
n
u
(1)
n
+ E
(1)
n
u
(0)
n
, (2.39)

2
: H
0
u
(2)
n
+ H
1
u
(1)
n
= E
(0)
n
u
(2)
n
+ E
(1)
n
u
(1)
n
+ E
(2)
n
u
(0)
n
. (2.40)
The zeroth-order equation (2.38) clearly is the same as (2.35). In order to derive the
rst-order correction, we multiply (2.39) by u
(0)
q
and integrate over D. Then, observing
that
(u
(0)
q
, u
(0)
n
) =
qn
,
29
we obtain
(u
(0)
q
, H
0
u
(1)
n
) + (u
(0)
q
, H
1
u
(0)
n
) = E
(0)
n
(u
(0)
q
, u
(1)
n
) + E
(1)
n

qn
.
Integrating by parts twice (or employing the self-adjointness of H
0
) it follows
(u
(0)
q
, H
0
u
(1)
n
) = (H
0
u
(0)
q
, u
(1)
n
) = E
(0)
q
(u
(0)
q
, u
(1)
n
),
and therefore,
(E
(0)
q
E
(0)
n
)(u
(0)
q
, u
(1)
n
) + (u
(0)
q
, H
1
u
(0)
n
) = E
(1)
n

qn
.
For q = n this gives an expression for E
(1)
n
only depending on (u
(0)
n
)
n
:
E
(1)
n
= (u
(0)
n
, H
1
u
(0)
n
). (2.41)
For q ,= n we have
(u
(0)
q
, u
(1)
n
) =
(u
(0)
q
, H
1
u
(0)
n
)
E
(0)
q
E
(0)
n
. (2.42)
This is possible since the non-degeneracy assumption implies E
(0)
q
,= E
(0)
n
for all q ,= n.
The sequence (u
(0)
n
) is an orthonormal basis, so we can develop u
(1)
n
in this basis:
u
(1)
n
=

q
(u
(0)
q
, u
(1)
n
)u
(0)
q
. (2.43)
In this sum we need an expression for the term q = n, (u
(0)
n
, u
(1)
n
). In fact, this term is not
determinable from the above calculation and can be chosen freely. We make the choice
(u
(0)
n
, u
(1)
n
) = 0. In view of (2.42), (2.43) becomes
u
(1)
n
=

q,=n
(u
(0)
q
, H
1
u
(0)
n
)
E
(0)
q
E
(0)
n
u
(0)
q
.
Thus, up to rst order, the eigenfunctions are given by
u
(0)
n
+ u
(1)
n
= u
(0)
n
+

q,=n
(u
(0)
q
, H
1
u
(0)
n
)
E
(0)
q
E
(0)
n
u
(0)
q
and the eigenvalues are
E
(0)
n
+ E
(1)
n
= E
(0)
n
+ (u
(0)
n
, H
1
u
(0)
n
).
Notice that these corrections only depend on the unperturbed eigenfunctions u
(0)
n
which
are assumed to be known.
In order to derive the second-order correction, we multiply (2.40) by u
(0)
q
and integrate
over D:
(u
(0)
q
, H
0
u
(2)
n
) + (u
(0)
q
, H
1
u
(1)
n
) = E
(0)
n
(u
(0)
q
, u
(2)
n
) +E
(1)
n
(u
(0)
q
, u
(1)
n
) + E
(2)
n

qn
.
30
As above, the rst term on the left-hand side equals
(u
(0)
q
, H
0
u
(2)
n
) = E
(0)
q
(u
(0)
q
, u
(2)
n
),
such that
(E
(0)
q
E
(0)
n
)(u
(0)
q
, u
(2)
n
) + (u
(0)
q
, H
1
u
(1)
n
) = E
(1)
n
(u
(0)
q
, u
(1)
n
) + E
(2)
n

qn
.
Using (2.41) and (2.43), the case q = n yields
E
(2)
n
= (u
(0)
n
, H
1
u
(1)
n
) E
(1)
n
(u
(0)
n
, u
(1)
n
)
=

q
(u
(0)
q
, u
(1)
n
) (u
(0)
n
, H
1
u
(0)
q
) (u
(0)
n
, H
1
u
(0)
n
) (u
(0)
n
, u
(1)
n
)
=

q,=n
(u
(0)
n
, H
1
u
(0)
q
) (u
(0)
q
, u
(1)
n
)
=

q,=n
(u
(0)
n
, H
1
u
(0)
q
)
(u
(0)
q
, H
1
u
(0)
n
)
E
(0)
q
E
(0)
n
.
In the last equation we employed (2.42). Thus, the second-order correction to the eigen-
values is
E
(0)
n
+ E
(1)
n
+
2
E
(2)
n
= E
(0)
n
+ (u
(0)
n
, H
1
u
(0)
n
) +

q,=n
(u
(0)
n
, H
1
u
(0)
q
)(u
(0)
q
, H
1
u
(0)
n
)
E
(0)
q
E
(0)
n
.
It remains to compute the scalar products. We write
(u
(0)
n
, H
1
u
(0)
q
) =
i
2
m
k (u
(0)
n
, u
(0)
q
) +

2
2m
[k[
2
(u
(0)
n
, u
(0)
q
)
=
i
2
m
k P
nq
+

2
2m
[k[
2

nq
,
where
P
nq
=
_
D
u
(0)
n
u
(0)
q
dx =
_
D
u
(0)
n
u
(0)
q
dx,
since u
(0)
n
is real. The periodicity of u
(0)
n
on D gives
P
nn
=
1
2
_
D
div[(u
(0)
n
)
2
] dx = 0,
and therefore,
(u
(0)
n
, H
1
u
(0)
q
) =
_

2
2m
[k[
2
: n = q

i
2
m
k P
nq
: n ,= q.
31
This shows that E
n
(k) is, up to second order in ,
E
n
(k) = E
(0)
n
+ E
(1)
n
+
2
E
(2)
n
= E
(0)
n
+

2
2m
[k[
2


4
m
2

q,=n
(k P
nq
)(k P
qn
)
E
(0)
q
E
(0)
n
(2.44)
= E
(0)
n
+

2
2m
[k[
2


4
m
2

q,=n

j,
k
j
k

P
nqj
P
qn
E
(0)
q
E
(0)
n
= E
(0)
n
+

2
2

j,
k
j
k

j
which proves the theorem.
Equation (2.44) shows that the rst-order correction of the energy yields simply the
free-electron mass. The second-order correction is needed to obtain an eective mass
which is dierent from the free-electron mass. This is the reason why we computed the
corrections up to second-order.
Theorem 2.11 can be applied to the bottom of the conduction band of most semi-
conductors since the eigenstates are non-degenerate (the energy bands do not cross; see
Figure 2.9). However, the top of the valence band in all semiconductors is degenerate (the
valence bands cross; see Figure 2.9) and hence, the above result does not hold. Mathe-
matically, we have in such a situation several eigenfunctions with the same eigenvalue, for
instance for n ,= q,
H
0
u
(0)
n
= E
(0)
n
u
(0)
n
, H
0
u
(0)
q
= E
(0)
q
u
(0)
q
but E
(0)
n
= E
(0)
q
.
Then, the expression (2.37) is not dened. It is still possible to derive a formula similar
to (2.36) in the degenerate case by applying degenerate perturbation theory. The idea is
to nd a linear combination
u
(0)
n
=
A

=1
c

u
(0)
n,
of the eigenfunctions u
(0)
n,
with the same energy E
(0)
n
such that the nominator in the
rst-order correction
u
(1)
n
=

q,=n
(u
(0)
q
, H
1
u
(0)
n
)
E
(0)
q
E
(0)
n
u
(0)
q
vanishes if E
(0)
q
= E
(0)
n
. The problem is to nd coecients c

such that ( u
(0)
q
, H
1
u
(0)
n
) = 0
[11, Sec. 4.2]. It can be shown that the energies E
n
(k) are, up to second order, the
eigenvalues of the matrix H
nk
R
AA
with elements [65, Sec. 4.1.4]
(H
nk
)

= E
(0)
n

i
2
m

j
k

nnj
+

2
2

j,
k
j
k

(m

j
,
32

P
j
nq
is dened similarly as in Theorem 2.11 and
m
(m

j
=
j

2
2
m

q,=n
A

=1

nqj

qn
E
(0)
q
E
(0)
n
. (2.45)
As a nal remark we notice that analogous results as above can be derived for holes
in the valence band. In this case, the energy E
n
(k) can be approximately written as
E
n
(k) = E
c
n
(0)

2
2
k

(m

h
)
1
k,
where E
c
n
(0) is the top energy of the valence band and m

h
is the eective mass tensor for
the holes, similarly dened as above. In this case,

P

nnj
= 0 for all j, such that the linear
term in k in (2.45) vanishes.
2.4 Semiconductor statistics
In this section we will answer the question how many electrons and holes are in a semi-
conductor which is in thermal equilibrium (i.e. no current ow)? With the mean number
of electrons in a quantum state of energy E, f(E), the answer is
N = 2

kB
f(E
n
(k)), (2.46)
where the factor 2 takes into account the two possible states of the spin of an electron
and B is the Brillouin zone. This leads to two questions:
How can the sum over many k can be computed practically?
How does the function f depend on the energy?
To answer the rst question, we make the following observation. Consider a chain of
N +1 atoms with distance d. Then the length of the chain is L = Nd. The Bloch function

k
(x) = e
ikx
u
k
(x) (see Section 2.1) satises at the chain boundaries

k
(0) =
k
(L)
and u
k
(x) is periodic, in particular, u
k
(0) = u
k
(L). Then
u
k
(0) =
k
(0) =
k
(L) = e
ikL
u
k
(L) = e
ikL
u
k
(0)
which implies that
kL = 2j, j = 0, . . . , N 1.
33
The index j runs from 0 to N 1 since the one-dimensional Brillouin zone equals here
k B = [0, 2/d) (see Lemma 2.4). Thus, the wave vector k can take one of the discrete
values
k
j
=
2j
L
=
2j
Nd
j = 0, . . . , N 1.
Typically, L = 1m = 10
6
m and d = 10
10
m, so N = 10
4
. Therefore, one can consider
k to be continuous. The sum
N1

j=0
g(k
j
)
for some function g then transforms to the integral
N1

j=0
g(k
j
) =
N1

j=0
g
_
2j
Nd
_

_
N
0
g
_
2j
Nd
_
dj =
L
2
_
2/d
0
g(k) dk.
In d space dimensions, the factor L/2 becomes (L/2)
d
. In the continuum limit N
and d 0 such that Nd is nite we can extend the integration to R
d
and write
vol()
(2)
d
_
R
d
g(k) dk instead of

kB
g(k), (2.47)
where vol() is the volume of the semiconductor. The advantage of the integral formula-
tion is that integrals can be more easily computed than sums.
The answer to the second question is contained in the following lemma.
Lemma 2.12 The mean number of electrons in a quantum state of energy E is given by
the Fermi-Dirac distribution function
f(E) =
1
1 + e
(Eq)/k
B
T
, (2.48)
where k
B
is the Boltzmann constant, T the (electron) temperature and the chemical
potential.
The two parameters T and can be considered here as given by the corresponding
quantum state. A deeper understanding is possible by means of thermodynamics, but
we refer to [11, Ch. 5] for details since the signicance of T and will become more
transparent in subsequent chapters (see also Remark 2.13).
For the formal proof of Lemma 2.12 we notice that electrons are fermions, i.e. particles
with half-integral spin, satisfying the following properties:
Electrons cannot be distinguised from each other.
The Pauli exclusion principle holds, i.e., each quantum state can be occupied by not
more than two electrons with opposite spins.
34
Proof of Lemma 2.12: We proceed as in [11, p. 293 .]. We arrange M electrons into N
bands each of which has g
n
quantum states, n = 1, . . . , N. Suppose that m
n
electrons
are occupying quantum states in the n-th band, where m
n
g
n
. Notice that, by the
Pauli exclusion principle, each electron is occupying exactly one quantum state. The
number of dierent arrangements in the n-th band Q
n
equals the number of all possible
congurations,
g
n
(g
n
1) . . . (g
n
m
n
+ 1),
divided by the number of all possible permutations of the m
n
electrons (since they are
indistinguishable),
m
n
!,
hence
Q
n
=
g
n
(g
n
1) . . . (g
n
m
n
+ 1)
m
n
!
=
g
n
!
(g
n
m
n
)!m
n
!
.
The total number of congurations reads in the limit N
Q(m
1
, m
2
, . . .) =

n=1
Q
n
=

n=1
g
n
!
(g
n
m
n
)!m
n
!
.
In order to manipulate this function, it is convenient to consider
ln Q(m
1
, m
2
, . . .) =

n=1
[ln g
n
! ln(g
n
m
n
)! ln m
n
!].
Using the Stirling formula
n!
n
n
e
n
or ln n! nln n n (n ),
we obtain approximately
ln Q(m
1
, m
2
, . . .) =

n=1
[ln g
n
! (g
n
m
n
) ln(g
n
m
n
) m
n
ln m
n
+g
n
].
This function is in some sense related to the thermodynamic entropy (see [11, p. 268]).
The most probable conguration of (m
1
, m
2
, . . .) is that one which maximizes ln Q, under
the constraints that the particle number and the energy are conserved:
max
m
j
ln Q(m
1
, m
2
, . . .) such that

n=1
m
n
= M,

n=1
E
n
m
n
= E.
We solve this constrained extremal problem with Lagrange multipliers, i.e., we extremize
F(
1
,
2
; m
1
, m
2
, . . .) = ln Q +
1
_

n=1
m
n
M
_
+
2
_

n=1
E
n
m
n
E
_
.
35
A necessary condition is
0 =
F
m
j
= ln(g
j
m
j
) + 1 (ln m
j
+ 1) +
1
+
2
E
j
= ln
_
g
j
m
j
1
_
+
1
+
2
E
j
.
Solving for m
j
/g
j
yields
m
j
g
j
=
1
1 + e

2
E
j
.
Dening the temperature T and the chemical potential by

1
=
q
k
B
T
and
2
=
1
k
B
T
,
we obtain
m
j
g
j
=
1
1 + e
(E
j
q)/k
B
T
= f(E
j
).
Since the left-hand side is the mean number of electrons in a quantum state of energy E
j
,
the lemma is shown.
Remark 2.13 The properties of the Fermi-Dirac distribution can be understood as fol-
lows (see also [11, p. 298 f.]). At zero temperature, this function becomes
f(E) =
_
1 for E < q
0 for E > q
and f(q) =
1
2
(see Figure 2.13). This means that all states which have an energy smaller than the
chemical potential are occupied, and all states with an energy larger than q are empty.
Physically, this behavior comes from the Pauli principle according to which two electrons
do not occupy the same quantum state. Consequently, at zero temperature, the states
with lowest energy are lled rst. The energy of the state lled by the last particle is
equal to the chemical potential. For non-zero temperature, there is a positive probability
that some energy states above q will be occupied, i.e., some particles jump to higher
energy levels due to thermal exertation.
We notice that the chemical potential q (more precisely, the product of elementary
charge and chemical potential) in semiconductors is referred to as the Fermi level or Fermi
potential and is usually denoted by E
F
.
For energies much larger than the Fermi energy E
F
= q in the sense of EE
F
k
B
T,
we can approximate the Fermi-Dirac distribution by the Maxwell-Boltzmann distribution
F(E) = e
(EE
F
)/k
B
T
since 1/(1+e
x
) e
x
for x 1 (Figure 2.13). Semiconductors whose electron distribution
can be described by this distribution are called non-degenerate. Materials in which the
Fermi-Dirac distribution has to be used are termed degenerate.
36
1
1
2
0
T > 0
T = 0
E
F
= q E
Maxwell-
Boltzmann
approximation
Fermi-Dirac
distribution
Figure 2.13: The Fermi-Dirac distribution at zero and non-zero temperature and the
Maxwell-Boltzmann approximation.
Now we wish to determine the density of states, i.e. the number of quantum states
(n, k) with energy E in the semiconductor :
g(E) =
2
vol()

kB
(E E
n
(k)), (2.49)
where again the factor 2 comes from the two spin states of an electron and (z) = 1 if
z = 0 and (z) = 0 if z ,= 0. We show:
Lemma 2.14 The density of states in the continuum limit in R
d
(d 1) reads as
g(E) =
2
(2)
d

n
_
R
d
(E E
n
(k)) dk, (2.50)
where is the delta distribution (see below). In the case d > 1 we can also write
g(E) =
2
(2)
d

n
_
E
1
n
(E)
dF
d1
[
k
E
n
(k)[
,
where dF
d1
is the element of the surface E
1
n
(E).
Before we can give a heuristic proof of this lemma, we need some preparations. First,
we explain the delta distribution. Heuristically, is dened by
(z) :=
_
: z = 0
0 : z ,= 0
and
_
R
(z) dz = 1. (2.51)
Clearly, there is no function satisfying these conditions. However, in the theory of distri-
butions it is possible to give a sense to (2.51). More precisely, is dened as a functional,
, ) = (0) for appropriate : R R.
37
Usually, the bracket is written as
_
R
(z)(z) dz = (0), (2.52)
which is motivated by (2.51). We refer to [22] for details about distributions.
The second tool needed for the proof of Lemma 2.14 is the coarea formula (see, e.g., [30,
46]).
Theorem 2.15 (Coarea formula)
Let f : B R
d
R (d > 1) be continuous and E : B R be continuously dierentiable
with
k
E(k) ,= 0 for all k B. Then
_
B
f(k) dk =
_
R
_
E
1
()
f(k)
dF
d1
(k)
[
k
E(k)[
d,
where dF
d1
(k) is the (d 1)-dimensional hypersurface element.
Proof: In the following we motivate the coarea formula for the case d = 3. The idea
of the proof is to make a transformation of the k- to the -variable. Let (u, v) be a
parametrization of the surface E
1
() = k : E(k) = R
3
(see Figure 2.14). Then
k = k(u, v, E) and = E(k(u, v, )). Taking partial derivatives gives
0 =

u
=
k
E(k)
k
u
,
0 =

v
=
k
E(k)
k
v
,
1 =

=
k
E(k)
k

.
The rst two equations imply that
k
E k/u and
k
E k/v and

k
E(k) |
_
k
u

k
v
_
. (2.53)
The third equation shows that [k/[ = 1/[
k
E[ and
k
E | k/. Then we infer from
(2.53)
k

|
_
k
u

k
v
_
. (2.54)
We conclude from the transformation theorem and (2.54) that
dk =

det
dk
d(u, v, )

d(u, v) d =

det
_
k
u
,
k
v
,
k

d(u, v)d
=


_
k
u

k
v
_

d(u, v)d =

k
u

k
v

d(u, v)d
=
1
[
k
E[
dFd
38
and hence,
_
B
f(k) dk =
_
R
_
E(k)=
f(k)
[
k
E[
dF
d1
d,
which motivates the theorem.
E(k) =
k

k
u
k
v
Figure 2.14: Illustration for the motivation of Theorem 2.15.
Proof of Lemma 2.14: In the continuum limit, the substitution (2.47) in the denition
(2.49) of g(E) immediately gives the rst equality in (2.50). The second one follows from
the denition (2.52) of the distribution and the coarea formula since for E > 0,
_
R
d
(E E
n
(k)) dk =
_
R
__
En(k)=
dF
d1
[
k
E
n
[
_
(E E
n
(k)) d
=
_
R
_
En(k)=
dF
d1
[
k
E
n
[
(E ) d
=
_
En(k)=E
dF
d1
[
k
E
n
[
.
The above lemmas allow us to determine the particle densities.
Lemma 2.16 The electron and hole densities are given by
n =
_
R
g
c
(E)f(E) dE, p =
_
R
g
v
(E)(1 f(E)) dE
where g
c
(E), g
v
(E) are the densities of states of the conduction or valence band, respec-
tively,
g
c
(E) =
2
(2)
d
_
R
d
(E E
c
(k)) dk, g
v
(E) =
2
(2)
d
_
R
d
(E E
v
(k)) dk,
and f(E) is the Fermi-Dirac distribution function (2.48).
39
Proof: The electron density is dened as the number of electrons N per volume vol().
Thus, by (2.46), (2.47) and (2.52), for a special band E
c
(k) and B = R
d
,
n =
N
vol()
=
2
vol()

kB
f(E
c
(k)) =
2
(2)
d
_
R
d
f(E
c
(k)) dk
=
2
(2)
d
_
R
d
_
R
(E E
c
(k))f(E) dE dk =
_
R
g
c
(E)f(E)dE.
In the last step we have used (2.50). The formula for p follows similarly, taking into
account that the mean number of holes in a quantum state of energy E equals the mean
number of empty states of energy E, 1 f(E).
In the parabolic band approximation, (2.31) and (2.32), the particle densities can be
computed more explicitely. For this, we rst compute the density of states.
Lemma 2.17 In the parabolic band approximation E
n
(k) = E
0
+(
2
/2m

)[k[
2
we obtain
for E E
0
:
g(E) =
m

2
_
2m

(E E
0
)

for three-dimensional carriers,


g(E) =
m

2
for two-dimensional carriers,
g(E) =
m

_
2m

(E E
0
)
for one-dimensional carriers.
For E < E
0
, we have g(E) = 0 in all three cases.
It is possible in modern quantum devices to conne carriers in one (or two) dimensions,
i.e., the carriers are conned in the x-y-plane (or in the x-direction) and are free to move
in the z-direction (or in the y-z-plane). Such structures can be constructed with so-
called semiconductor heterostructures and are called quantum wires or quantum wells,
respectively (see [45, Sec. 1.5.2]).
Proof: We start from the rst equality in (2.50), use spherical coordinates (, , ) in the
three-dimensional case and substitute z =
2

2
/2m

:
g(E) =
2
(2)
3
_
R
3

_
E E
0


2
2m

[k[
2
_
dk
=
1
4
3
_
2
0
_

0
_

0

_
E E
0


2
2m

2
_

2
sin d d d
=
4
4
3
m

2m

_

0
(E E
0
z)

z dz.
40
Introducing the Heaviside function H by H(x) = 0 for x < 0 and H(x) = 1 for x > 0 we
obtain from (2.52):
g(E) =
m

2m

_
R
(E E
0
z)

zH(z) dz =
m

2m

_
E E
0
H(E E
0
).
For the two-dimensional case, we start again from (2.50) and use polar coordinates
(, ) and the substitution z =
2

2
/2m

,
g(E) =
2
(2)
2
_
2
0
_

0

_
E E
0


2
2m

2
_
d d
=
2
2
2
m

2
_

0
(E E
0
z) 1 dz =
m

2
.
Finally, for the one-dimensional case,
g(E) =
2
(2)
_
R

_
E E
0


2
2m

k
2
_
dk =
1

2m

2
_

0
(E E
0
z)
dz

z
=
m

_
2m

(E E
0
)
.

Remark 2.18 In the non-parabolic band approximation
E
n
(1 +E
n
) =

2
2m

[k[
2
, > 0
(see (2.30)), the densities of states become in the three-dimensional case
g(E) =
m

2m

1 + E (1 + 2E)
and in the two-dimensional case
g(E) =
m

2
(1 + 2E)
(see [45, Problem 1.4]).
Lemma 2.19 Let the conduction and valence bands be given by the parabolic band ap-
proximations (2.31) and (2.32). Then, for three-dimensional particles,
n = N
c
F
1/2
_
q E
c
k
B
T
_
, p = N
v
F
1/2
_
E
v
q
k
B
T
_
,
where
N
c
= 2
_
m

e
k
B
T
2
2
_
3/2
, N
v
= 2
_
m

h
k
B
T
2
2
_
3/2
(2.55)
are the eective densities of states and
F
1/2
(z) =
2

_

0

xdx
1 + e
xz
, z R,
is the Fermi integral (of index 1/2). Furthermore, m

e
and m

h
denote the (isotropic)
eective mass of the electrons and holes, respectively.
41
Proof: From Lemmas 2.16 and 2.17 and the substitution x = (E E
c
)/k
B
T we obtain
n =
m

2m

_

Ec

E E
c
1 + e
(Eq)/k
B
T
dE =
4

_
m

e
k
B
T
2
2
_
3/2
_

0

xdx
1 + e
x(qEc)/k
B
T
= N
c
F
1/2
_
q E
c
k
B
T
_
.
In a similar way,
p =
m

2
_
2m

_
Ev

_
E
v
E
e
(Eq)/k
B
T
1 + e
(Eq)/k
B
T
dE
=
4

_
m

h
2
2
_
3/2
_
Ev

E
v
E dE
1 + e
(Eq)/k
B
T
=
4

_
m

h
k
B
T
2
2
_
3/2
_
0

xdx
1 + e
x(Evq)/k
B
T
= N
v
F
1/2
_
E
v
q
k
B
T
_
.

We compute the carrier densities in some special situations.


Lemma 2.20 The electron density in a quantum well equals
n =
m

e
k
B
T

2
ln
_
1 + e
(qEc)/k
B
T
_
.
Proof: In a quantum well, electrons are conned in one direction. Therefore, using the
density of states function for two-dimensional carriers (see Lemma 2.17),
n =
m

2
_

Ec
dE
1 + e
(Eq)/k
B
T
=
m

e
k
B
T

2
_
ln(1 +e
(Eq)/k
B
T
)

Ec
=
m

e
k
B
T

2
ln(1 +e
(Eq)/k
B
T
).

Lemma 2.21 The electron and hole densities in the three-dimensional parabolic band and
Maxwell-Boltzmann approximation are
n = N
c
exp
_
q E
c
k
B
T
_
, p = N
v
exp
_
E
v
q
k
B
T
_
,
where N
c
and N
v
are the eective densities of states dened in (2.55).
42
Proof: For z we can approximate
F
1/2
(z) = e
z
2

_

0

x dx
e
z
+ e
x
e
z
2

_

0

xdx
e
x
= e
z
2

_
3
2
_
= e
z
,
where (p) is the function,
(p) =
_

0
x
p1
e
x
dx,
with the properties (
1
2
) =

and (p+1) = p(p), p > 0. Thus, the result follows from


Lemma 2.19.
Finally, we discuss two notions needed in the subsequent chapters, the intrinsic density
and the doping of semiconductors.
A pure semiconductor with no impurities is called an intrinsic semiconductor. In this
case, electrons in the conduction band can only come from valence band levels leaving
a vacancy behind them. Vacancies in the valence band are called holes (Figure 2.11).
Therefore, the number of electrons in the conduction band is equal to the number of holes
in the valence band,
n = p = n
i
.
The quantity n
i
is called intrinsic density. It can be computed in the non-degenerate
parabolic band case from Lemma 2.21:
n
i
=

np =
_
N
c
N
v
exp
_
E
v
E
c
2k
B
T
_
=
_
N
c
N
v
exp
_

E
g
2k
B
T
_
(2.56)
since the energy gap is E
g
= E
c
E
v
. This allows to determine the Fermi energy E
F
= q
of an intrinsic semiconductor [3, (28.22)]:
E
F
= E
c
+ k
B
T ln
n
N
c
= E
c
+ k
B
T ln
n
i
N
c
= E
c

E
g
2
+
k
B
T
2
ln
N
v
N
c
=
1
2
(E
c
+E
v
) +
3
4
k
B
T ln
m

h
m

e
.
This asserts that at zero temperature, the Fermi energy lies precisely in the middle of
the energy gap. Furthermore, since ln(m

h
/m

e
) is of order one, the correction is only of
order k
B
T for non-zero temperature. In most semiconductors at room temperature, the
energy gap is much larger than k
B
T 0.0259 eV (T = 300K). This shows that the
non-degeneracy assumptions
E E
F
E
c
E
F
=
E
g
2
+
3
4
k
B
T ln
m

h
m

e
k
B
T,
E
F
E E
F
E
v
=
E
g
2
+
3
4
k
B
T ln
m

h
m

e
k
B
T
43
E
c
E
v
E
g
k
B
T
E
E
v
+
Eg
2
E
F
k
Figure 2.15: Illustration of the energy gap E
g
in relation to the energies E
c
, E
v
, and E
F
(see [3, Fig. 28.10]).
are satised and that the result is consistent with our assumptions (see Figure 2.15).
The intrinsic density is too small to result in a signicant conductivity for non-zero
temperature. For instance, in silicon we have n
i
6.93 10
9
cm
3
compared to N
c
, N
v

10
19
cm
3
. Replacing some atoms in the semiconductor crystal by atoms which provide free
electrons in the conduction band or free holes in the valence band allows to increase the
conductivity. Such a process is called doping. Impurities are called donors if they supply
additional electrons to the conduction band, and acceptors if they supply additional holes
to (i.e. capture electrons from) the valence band. A semiconductor which is doped with
donors is termed n-type semiconductor, and a semiconductor doped with acceptors is called
p-type semiconductor. For instance, when we dope a germanicum crystal, whose atoms
have each 4 valence electrons, with arsenic, which has 5 valence electrons per atom, each
arsenic atom provides one additional electron (see Figure 2.16). These additional electrons
are only weakly bound to the arsenic atom. Indeed, the binding energy is about 0.013 eV
(see [3, Table 28.2]) which is much smaller than the thermal energy k
B
T 0.026 eV at
room temperature. More generally, denoting by E
d
and E
a
the energies of a donor electron
and an acceptor hole, respectively, then E
c
E
d
and E
a
E
v
are small compared to k
B
T
(see Figure 2.17). This means that the additional carriers contribute at room temperature
to the electron and hole density and increase the conductivity of the semiconductor.
Let N
D
(x), N
A
(x) denote the densities of the donor and acceptor impurities, respec-
tively. Then the doping prole or doping concentration is C(x) = N
A
(x) N
D
(x) and the
total space charge (see Chapter 1) is given by
= qn + qp + qN
A
(x) qN
D
(x) = q(n p C(x)).
44
Ge Ge Ge
Ge Ge As
Figure 2.16: Germanium crystal with an arsenic impurity. The bullets represent electrons
bound to the atoms. The arsenic atom provides an additional electron which is not bound.

`
k
E
v
E
c
E
E
a
E
d
`

E
g
Figure 2.17: Illustration of the donor and acceptor energy levels E
d
and E
a
.
45
3 Classical Kinetic Models
3.1 The Liouville equation
We rst analyze the motion of M particles with mass m in a vacuum under the action
of a force. The particles are described as classical particles, i.e., we associate the position
vector x
i
R
d
and the velocity vector v
i
R
d
with the i-th particle of the ensemble.
The trajectories (x
i
(t), v
i
(t)) of the particles satisfy Newtons equations in the (2d M)-
dimensional ensemble position-velocity space
x = v, v =
1
m
F(t), t > 0, (3.1)
with initial conditions
x(0) = x
0
, v(0) = v
0
. (3.2)
Here, x = (x
1
, . . . , x
M
), v = (v
1
, . . . , v
M
), F = (F
1
, . . . , F
M
) is a force and the dot means
dierentiation with respect to t. For instance, the forces can be given by the electric eld
acting on the electron ensemble:
F
i
= qE(x, t), i = 1, . . . , M,
where q > 0 is the elementary charge. In this case the forces are independent of the ve-
locity. In semiconductors, M is a very large number (typically, M 10
4
) and therefore,
the numerical solution of (3.1)(3.2) is very expensive. It seems reasonable to use a sta-
tistical description. Instead of the precise initial conditions x(0) and v(0) we are given
the probability density f
I
(x, v) of the initial position and velocity of the particles. The
integral
_
A
f
I
(x, v) dxdv
is the probability to nd the particle ensemble at time t = 0 in the subset A of the
(v, x)-phase space.
Let f(x, v, t) be the probability density of the ensemble at time t. We wish to derive a
dierential equation for f. In Chapter 1 we have already argued that under the assumption
that f is constant along the trajectories (x(t), v(t)), i.e.
f(x(t), v(t), t) = f
I
(x
0
, v
0
), (3.3)
dierentiation with respect to t leads to the dierential equation
0 =
d
dt
f(x(t), v(t), t) =
t
f + x
x
f + v
v
f
=
t
f + v
x
f +
1
m
F
v
f. (3.4)
46
This equation is referred to as the (classical) Liouville equation. It has to be solved in
R
2dM
R and is supplemented by the initial condition
f(x, v, 0) = f
I
(x, v), (x, v) R
2dM
.
In the following we show under which conditions (3.3) is satised.
Theorem 3.1 (Liouville)
Assume that
M

j=1
_
x
i
x
i
+
v
i
v
i
_
= 0 for all t > 0. (3.5)
Then (3.3) holds.
Notice that (3.5) is trivially satised if Newtons laws (3.1) are assumed. However,
in the semiconductor case, (3.1) has to be replaced by the semi-classical picture (2.33)
for which (3.5) will be veried later. For details about the Liouville equation for gases,
see [19].
t = 0
dv
0
dx
0
dv
dx
t > 0
(x(t), v(t))
v
x
Figure 3.1: Innitesimal volume elements dx
0
dv
0
at time t = 0 and dxdv at time t > 0 in
the phase space.
Proof: We will examine how the innitesimal phase-space element dx
0
dv
0
will change
during time. In particular, we show that [23, Sec. 1.2]
dxdv(t) = dx
0
dv
0
. (3.6)
This implies the assertion (3.3). Indeed, the number of particles at time t in the volume
dxdv(t) is given by
f(x(t), v(t), t)dxdv(t).
47
The conservation of particles implies that
f(x(t), v(t), t)dxdv(t) = f(x
0
, v
0
, 0)dx
0
dv
0
and therefore, (3.6) gives (3.3).
In the following we show (3.6). Assuming sucient regularity of the functions, the
mapping (x
0
, v
0
) (x(t), v(t)) is continuously dierentiable and invertible with continu-
ously dierentiable inverse. Then
dxdv(t) = J(t)dx
0
dv
0
, (3.7)
where
J(t) =

det
(x(t), v(t))
(x
0
, v
0
)

det
_
_
x(t)
x
0
x(t)
v
0
v(t)
x
0
v(t)
v
0
_
_

.
We claim that (3.5) implies
t
J(t) = 0. In order to see this, we use the Taylor expansion
x(t +) = x(t) + x(t) + O(
2
), v(t + ) = v(t) + v(t) +O(
2
).
This gives
x(t + )
x(t)
= I +
x(t)
x(t)
+ O(
2
),
v(t + )
v(t)
= I +
v(t)
v(t)
+ O(
2
),
where I denotes the unit matrix in (dM) (dM). Since x(t) and v(t) are independent
variables, we have
x(t)
v(t)
= 0,
v(t)
x(t)
= 0
and therefore,
x(t + )
v(t)
=
x(t)
v(t)
+ O(
2
),
v(t + )
x(t)
=
v(t)
x(t)
+ O(
2
).
Thus we can compute
det
(x, v)(t + )
(x, v)(t)
= det
_
_
_
_
x(t + )
x(t)
x(t + )
v(t)
v(t +)
x(t)
v(t + )
v(t)
_
_
_
_
= det
_
_
_
_
I +
x(t)
x(t)

x(t)
v(t)

v(t)
x(t)
I +
v(t)
v(t)
_
_
_
_
+ O(
2
)
= 1 +

i
_
x
i
(t)
x
i
(t)
+
v
i
(t)
v
i
(t)
_
+ O(
2
)
= 1 +O(
2
),
48
by using assumption (3.5). Hence, by the chain rule,
J(t + ) =

det
_
(x, v)(t +)
(v, x)(t)
(v, x)(t)
(x
0
, v
0
)
_

det
(x, v)(t +)
(v, x)(t)
det
(v, x)(t)
(x
0
, v
0
)

det
(v, x)(t)
(x
0
, v
0
)

+ O(
2
) = J(t) + O(
2
),
and we obtain nally

t
J(t) = lim
0
1

(J(t + ) J(t)) lim


0
O() = 0.
The mapping (x
0
, v
0
) (x(t), v(t)) is the identity for t = 0 such that J(0) = 1. This
implies that J(t) = 1 for all t 0 and hence, (3.7) reduces to (3.6). The theorem is
proved.
The Liouville equation (3.4) is valid for an ensemble of particles moving in a vacuum
according to the laws of classical mechanics. An ensemble of M electrons moving in a
semiconductor crystal can be described semi-classically by the equations
x
j
= v
n
(k) =
1

k
j
E
n
(k),

k
j
=
q

x
j
V (x, t), j = 1, . . . , M (3.8)
(see (2.33)), where E
n
(k
j
) is the energy of the n-th band depending on the (pseudo-) wave
vector k and V (x, t) is the electrostatic potential. We denote as above x = (x
1
, . . . , x
M
)
and k = (k
1
, . . . , k
M
) R
dM
. We assume that the ensemble stays in the same energy band
so that we can drop the index n in the equations (3.8).
We claim that Liouvilles Theorem 3.1 can be applied to the semi-classical picture.
Since the classical momentum p = mv translates into the crystal momentum p = k, we
have to verify

j
_
x
i
x
i
+

k
i
k
i
_
= 0.
Actually, x depends on k but not on x, and

k depends on x but not on k (see (3.8)) such
that this condition is satised. Therefore, (3.3) holds and the Liouville equation (3.4) is
valid here. Observing that Newtons law (3.1) can be written as
p = F,
and that (3.8) can be reformulated as
p = q
x
V,
the Liouville equation (3.4) becomes

t
f +
1

k
E(k)
x
f
q

x
V
k
f = 0. (3.9)
49
Here, we have used that the classical term
p
=
1
m

v
translates into the semi-classical
expression
p
=
1

k
. Equation (3.9) is referred to as the semi-classical Liouville equation.
More precisely, the scalar products are dened as

k
E(k)
x
f =
M

j=1

k
j
E(k
j
)
x
j
f,
x
V
k
f =
M

j=1

x
j
V (x)
k
j
f.
Equation (3.9) has to be solved for x R
dM
, k B
M
, t > 0, where B is the Brillouin zone
(see Section 2.1). As the Brillouin zone is a bounded subset, we have to impose boundary
conditions. Often, periodic boundary conditions
f(x, k
1
, . . . , k
j
, . . . , k
M
, t) = f(x, k
1
, . . . , k
j
, . . . , k
M
, t), k
j
B,
for all j = 1, . . . , M are chosen [49, (1.2.47)]. This formulation makes sense since B is
point symmetric to the origin, i.e., k B if and only if k B. We also impose the
initial conditions
f(x, k, 0) = f
I
(x, k), x R
dM
, k B
M
.
Before we show some properties of the solutions to (3.9), we introduce the electron-
ensemble particle density and the electron-ensemble current density:
n(x, t) =
_
B
M
f(x, k, t) dk, (3.10)
J(x, t) = q
_
B
M
f(x, k, t)v(k) dk,
where v(k) = (v(k
1
), . . . , v(k
M
)).
Remark 3.2 Denition (3.10) of the number density is slightly inexact. Indeed, similar
to Lemma 2.16 and arguing as in [23, Sec. 1.2.2], the total number of electrons at time t
in the volume dxdk of the (conduction band) phase space is given by
dN(x, k) = f(x, k, t)g(x, k) dxdk,
where g(x, k) is the momentum density of states (in the conduction band) and f(x, k, t)
can be interpreted as an occupation distribution, i.e, the ratio of the number of occupied
states in the phase-space volume dxdk and the total number of quantum states in this
volume (in the conduction band). The volume in phase space occupied by a single quantum
state is of the order of (2)
d
. Thus, we have dxdk/(2)
d
quantum states in the volume
dxdk. To be precise, we have to take into account the spin of the electrons. Each quantum
state can be occupied by two electrons with opposite spin. Therefore, the number of
50
quantum states is twice such that g(x, k) = 2/(2)
d
. The correct denition of the electron
density becomes
n(x, t) =
_
B
M
f(x, k, t)g(x, k) dk =
_
B
M
f(x, k, t)
2
(2)
d
dk
(see [23] for details). In the following we ignore the factor 2/(2)
d
which can be achieved,
for instance, by scaling.
Lemma 3.3 The solution of the semi-classical Liouville equation (3.9) satises formally
the following properties:
(1) If f
I
(x, k) 0 for all x R
dM
, k B
M
, then f(x, k, t) 0 for all x R
dM
,
k B
M
.
(2) The following conservation law holds:

t
n
1
q
div
x
J = 0, x R
dM
, t > 0.
(3) The number of particles is conserved:
_
R
dM
n(x, t) dx =
_
R
dM
_
B
M
f
I
(x, k) dk dx.
Proof: The rst property follows immediately from (3.3). By formally integrating (3.9)
over k B
M
we obtain from the divergence theorem,

t
n =
_
B
M

t
f dk =
_
B
M
div
x
(v(k)f) dk +
q

_
B
M
div
k
(
x
V f) dk
=
1
q
div
x
J.
Finally, the last property follows from the second one after integrating over x R
dM
.
We notice that in the parabolic band approximation
E(k
j
) =

2
2m

[k
j
[
2
, j = 1, . . . , M, k
j
R
d
,
the semi-classical Liouville equation (3.9) reduces to its classical counterpart (3.4) since
v(k
j
) = k
j
/m

= p
j
/m

, which is the classical momentum-velocity relation, and


k
f =
(/m

)
v
f. We obtain from (3.9)

t
f + v
x
f
q
m

x
V
v
f = 0, x, v R
dM
, t > 0.
51
3.2 The Vlasov equation
The main disadvantage of the (semi-)clasical Liouville equation is that it has to be solved
in the very high-dimensional phase space R
2dM
. Typically, M 10
4
, d = 3, and the
dimension becomes 6 10
4
which is prohibitive for numerical simulations. In this section
we will formally derive a lower-dimensional equation, the so-called Vlasov equation. We
proceed as in [49, Sec. 1.3].
The idea of the derivation is rst to assume a certain structure of the force eld, then
to integrate the Liouville equation in a certain sub-phase space and nally to carry out the
formal limit M , where M is the number of particles. More precisely, we consider an
ensemble of M electrons and denote by x = (x
1
. . . , x
M
) R
dM
, k = (k
1
, . . . , k
m
) B
M
the position and wave vector coordinates of the particles, respectively. We impose the
following assumptions:
(1) The motion is governed by an external electric eld and by two-particle (long-range)
interaction forces
F
j
(x, t) = qE
ext
(x
j
, t) q
M

=1,j,=
E
int
(x
j
, x

), j = 1, . . . , M,
such that
E
int
(x
j
, x

) = E
int
(x

, x
j
). (3.11)
(2) The interaction force [E
int
[ is of order 1/M.
(3) The initial density is independent of the numbering of the particles:
f
I
(x
1
, . . . , x
M
, k
1
, . . . , k
M
) = f
I
(x
(1)
, . . . , x
(M)
, k
(1)
, . . . , k
(M)
) (3.12)
for all x
j
R
d
, k
j
B, j = 1, . . . , M, and for all permutations of 1, . . . , M.
(4) The sub-ensemble initial density
f
(a)
I
(x
1
, . . . , x
a
, k
1
, . . . , k
a
) =
_
(R
d
B)
Ma
f
I
dx
a+1
dx
M
dk
a+1
dk
M
can be factorized:
f
(a)
I
=
a

j=1
P
I
(x
j
, k
j
), a = 1, . . . , M 1.
We discuss the above assumptions. The rst assumption means that magnetic elds
are ignored. In fact, this hypothesis can be discarded, see [49, p. 26]. This assumption,
which is crucial for the derivation of the Vlasov equation, also means that the force eld F
i
52
exerted on the i-th electron is given by the sum of an external electric eld acting on the i-
th electron and of the sum of M1 two-particle interaction forces between the i-th electron
and all other electrons. The interaction force E
int
is independent of the electron indices
which interprets the fact that the electrons are indistinguishable. The action-reaction law
implies that the force exerted by the j-th electron on the -th electron is the negative force
of the -th electron on the j-th electron, i.e. E
int
(x
j
, x

) = E
int
(x

, x
j
). This property and
assumption (3) imply that also f(x, k, t) is independent of the numbering of the particle!
s for all t > 0. Finally, assumption (4) is needed for the limit M (in order to get a
nite force F
j
) and its meaning will become clear later.
We introduce the density f
(a)
of a subensemble consisting of a < M electrons:
f
(a)
(x
1
, . . . , x
a
, k
1
, . . . , k
a
, t) =
_
(R
d
B)
Ma
f(x, k, t) dx
(a+1)
dk
(a+1)
,
where dx
(a+1)
= dx
a+1
dx
M
and dk
(a+1)
= dk
a+1
dk
M
.
Theorem 3.4 Let the above assumptions (1)-(4) hold. Then the function f(x, k, t) is a
particular solution to the semi-classical Liouville equation (3.9) if M 1,
f
(a)
(x
1
, . . . , x
a
, k
1
, . . . , k
a
, t) =
a

j=1
P(x
j
, k
j
, t) (3.13)
and F(x, k, t) := MP(x, k, t) is a solution to the semi-classical Vlasov equation

t
F +v(k)
x
F
q

E
e

k
F = 0, x R
d
, k B, t > 0, (3.14)
F(x, k, 0) = F
I
(x, k), x R
d
, k B,
where
E
e
(x, t) = E
ext
(x, t) +
_
R
d
B
F(x
2
, k
2
, t)E
int
(x, x
2
) dx
2
dk
2
= E
ext
(x, t) +
_
R
d
n(x

, t)E
int
(x, x

) dx

(3.15)
and F
I
(x, k) = MP
I
(x, k) (see assumption (4)). Furthermore,
n(x, t) =
_
B
F(x, k, t) dk
represents the electron density, and we impose periodic boundary conditions:
F(x, k, t) = F(x, k, t), x R
d
, k B, t > 0.
53
Proof: We integrate the semi-classical Liouville equation

t
f +
M

j=1
v(k
j
)
x
j
f
q

j=1
E
ext
(x
j
, t)
k
j
f
q

j,=1
E
int
(x
j
, x

)
k
j
f = 0 (3.16)
with respect to x
a+1
, . . . , x
M
, k
a+1
, . . . , k
M
in order to obtain an equation for f
(a)
. We
reformulate these integrals term by term.
Clearly, the rst term on the left-hand side of (3.16) equals
t
f
(a)
after the integration.
For the second term we compute, using the divergence theorem,
M

j=1
_
(R
d
B)
Ma
v(k
j
)
x
j
f dx
(a+1)
dk
(a+1)
=
a

j=1
v(k
j
)
x
j
_
(R
d
B)
Ma
f dx
(a+1)
dk
(a+1)
+
M

j=a+1
_
(R
d
B)
Ma
div
x
j
(v(k
j
)f) dx
(a+1)
dk
(a+1)
=
a

j=1
v(k
j
)
x
j
f
(a)
.
Similarly,
M

j=1
_
(R
d
B)
Ma
div
k
j
(E
ext
(x
j
, t)f) dx
(a+1)
dk
(a+1)
=
a

j=1
div
k
j
_
E
ext
(x
j
, t)f
(a)
_
.
The last integral on the left-hand side becomes
M

j,=1
_
(R
d
B)
Ma
div
k
j
(E
int
(x
j
, x

)f) dx
(a+1)
dk
(a+1)
=
a

j,=1
E
int
(x
j
, x

)
k
j
f
(a)
+
M

j=a+1
M

=1
_
(R
d
B)
Ma
div
k
j
(E
int
f) dx
(a+1)
dk
(a+1)
+
a

j=1
M

=a+1
_
(R
d
B)
Ma
div
k
j
(E
int
(x
j
, x

)f) dx
(a+1)
dk
(a+1)
.
The second integral on the right-hand side of the above equation vanishes by the diver-
gence theorem. For the last integral we use the assumptions (3.11) and (3.12). Indeed, it
is possible to renumber the particles such that the last integral equals
a

j=1
(M a)
_
(R
d
B)
Ma
div
k
j
(E
int
(x
j
, x
a+1
)f) dx
(a+1)
dk
(a+1)
=
a

j=1
(M a)div
k
j
_
R
d
B
E
int
(x
j
, x

)f
(a+1)

dx

dk

,
54
where
f
(a+1)

= f
(a+1)
(x
1
, . . . , x
a
, x

, k
1
, . . . , k
a
, k

, t).
Thus, integration of (3.16) yields the system of equations
0 =
t
f
(a)
+
a

j=1
v(k
j
)
x
j
f
(a)

j=1
E
ext
(x
j
, t)
k
j
f
(a)

j,=1
E
int
(x
j
, x

)
k
j
f
(a)

j=1
div
k
j
_
R
d
B
(M a)E
int
(x
j
, x

)f
(a+1)

dx

dk

, (3.17)
where 1 a M 1. These equations are called the BBGKY hierarchy (from Bogoli-
ubov [12], Born and Green [14], Kirkwood [39], and Yvon [67]). By assumption (2), [E
int
[
is of the order of 1/M such that for M 1, the fourth term on the right-hand side of
(3.17) can be neglected. The term (Ma)E
int
however, stays nite and is approximately
equal to ME
int
. Therefore, we obtain for M 1
0 =
t
f
(a)
+
a

j=1
v(k
j
)
x
j
f
q

j=1
E
ext
(x
j
, t)
k
j
f
(a)

j=1
div
k
j
_
R
d
B
Mf
(a+1)

E
int
(x
j
, x

)dx

dk

. (3.18)
Now we claim that this equation is solved by the ansatz (3.13). In order to see this we
multiply (3.14) for (x, k) = (x
j
, k
j
) by
Q
j
=
1
M

,=j
P(x

, k

, t)
and take the sum for j = 1, . . . , a. Then, for F
j
= MP(x
j
, k
j
, t),
a

j=1
Q
j

t
F
j
=
a

j=1

,=j
P(x

, k

, t)
t
P(x
j
, k
j
, t) =
t
a

j=1
P(x
j
, k
j
, t) =
t
f
(a)
.
In a similar way,
a

j=1
Q
j
v(k
j
)
x
j
F
j
=
a

j=1
v(k
j
)
x
j
f
(a)
,
a

j=1
Q
j
E
ext
(x
j
, t)
k
j
F
j
=
a

j=1
E
ext
(x
j
, t)
k
j
f
(a)
.
55
a

j=1
Q
j
__
R
d
B
MP(x

, k

, t)E
int
(x
j
, x

) dx

dk

_

k
j
F
j
=
a

j=1
div
k
j
_
R
d
B
M
a

=1
P(x

, k

, t)P(x

, k

, t)E
int
(x
j
, x

) dx

dk

=
a

j=1
div
k
j
_
R
d
B
Mf
(a+1)
()
E
int
(x
j
, x

) dx

dk

.
Putting together the above computations, we see that the ansatz (3.13) indeed solves
(3.18). This proves the theorem.
The semi-classical Vlasov equation has the form of a Liouville equation for a single
particle with the force qE
e
. Many-particle physics is taken into account through the
eective eld E
e
which depends on the density n and hence on F. Thus, (3.14) is a
nonlinear equation with a nonlocal quadratic nonlinearity. The Vlasov equation describes
the macroscopic motion of many-particle systems with weak long-range forces. However, it
does not provide a description of strong short-range forces such as scattering of particles.
This case will be considered in Section 3.3.
As for the Liouville equation, the quantity F(x, k, t) can be interpreted as the proba-
bility density of a particle to be in the state (x, k) at time t. Indeed, we obtain from the
trajectory equations
x = v(k),

k =
q

E
e
, t > 0, x(0) = x
0
, k(0) = k
0
,
and (3.14) the equation
0 =
t
F + x
x
F +

k
k
F =
d
dt
F(x(t), v(t), t)
and thus
F(x(t), k(t), t) = F
I
(x
0
, k
0
) 0 for all t 0.
Finally, we wish to reformulate the nonlinear system (3.14)(3.15) in the case of the
Coulomb force in R
3
,
E
int
(x, y) =
q
4
s
x y
[x y[
3
, x, y R
3
, x ,= y, (3.19)
which is the most important long-range force between two electrons. Here, the permittivity

s
is a material constant. We assume that the external eld is generated by doping atoms
in the semiconductor crystal of charge +q:
E
ext
(x, t) =
+q
4
s
_
R
3
C(y)
x y
[x y[
3
dy, (3.20)
where C(x) is the doping concentration (see Section 2.4).
56
Proposition 3.5 In the case of the Coulomb force (3.19) and (3.20) the semi-classical
Vlasov equation (3.14)(3.15) can be written as the Vlasov-Poisson system

t
F + v(k)
x
F
q

x
V
e

k
F = 0, (3.21)

x
V
e
= q(n C), x R
3
, k B, t > 0. (3.22)
Proof: It is well known that the function
(x) =
1
4
_
R
3
f(y)
[x y[
dy, x R
3
,
solves the Poisson equation = f in R
3
under some regularity assumptions on f.
Therefore,
f(x) = (x) =
1
4
_
R
3
f(y)div
x
x y
[x y[
3
dy,
0 = curl (x) =
1
4
curl
x
_
R
3
f(y)
x y
[x y[
3
dy.
This shows that
divE
e
(x, t) = divE
ext
(x, t) +
_
R
3
n(x

, t)divE
int
(x, x

) dx

= divE
ext
(x, t)
q

s
n(x, t),
curl E
e
(x, t) = curl E
ext
(x, t) +
_
R
3
n(x

, t)curl E
int
(x, x

) dx

= curl E
ext
(x, t), x R
3
, t > 0.
Furthermore,
divE
ext
=
q

s
C, curl E
ext
= 0 in R
3
,
and hence,
divE
e
=
q

s
(n C), curl E
e
= 0 in R
3
.
Since E
e
is vortex-free, there exists a potential V
e
such that E
e
= V
e
. Thus,

s
V
e
=
s
divE
e
= q(n C),
and the proposition follows.
57
3.3 The Boltzmann equation
The Vlasov equation does not account for short-range particle interactions, like collisions
of the particles with other particles or with the crystal lattice. We wish to extend the
Vlasov equation to include scattering mechanisms which leads to the Boltzmann equation.
We present only a phenomenological derivation; for details on more rigorous derivations,
we refer to [4, Sec. 1.5.3], [15, Ch. 7], and [19]. The Boltzmann equation has been rst
formulated by Boltzmann in 1872 for the non-equilibrium transport of dilute gases [13].
The Vlasov equation along trajectories
dF
dt
= 0
states that the probability F (of occupation of states) does not change in time. Scattering
allows particles to jump to another trajectory. Our main assumption is that the rate of
change of F due to convection and the eective eld, dF/dt, and the rate of change of F
due to collisions, Q(F), balance:
dF
dt
= Q(F).
Clearly, this equation has to be understood along trajectories. By (3.4) this equation
equals (writing f instead of F)

t
f + v(k)
x
f
q

E
e

k
f = Q(f), x R
d
, k B, t > 0, (3.23)
where the eective eld is given by (3.15).
It remains to derive an expression for Q(f). We assume that scattering of particles
occurs instantaneously and only changes the crystal momentum of the particles. The rate
P(x, k
t
k, t) at which a particle at (x, t) changes its Bloch state k
t
into another Bloch
state k due to a scattering event is proportional to
the occupation probability f(x, k
t
, t) and
the probability 1 f(x, k, t) that the state (x, k) is not occupied at time t.
Here, we used the Pauli exclusion principle. Thus,
P(x, k
t
k, t) = s(x, k
t
, k)f(x, k
t
, t)(1 f(x, k, t)),
where the proportionality constant s(x, k
t
, k) is called the scattering rate. Then the rate
of change of f due to collisions is the sum of all in-scattering rates from k
t
to k minus the
out-scattering rate from k to k
t
,
P(x, k
t
k, t) P(x, k k
t
, t),
58
for all possible Bloch states k
t
in the volume element dk
t
. In the continuum limit, the sum
becomes an integral and we obtain
(Q(f))(x, k, t) =
_
B
[P(x, k
t
k, t) P(x, k k
t
, t)]dk
t
(3.24)
=
_
B
[s(x, k
t
, k)f
t
(1 f) s(x, k, k
t
)f(1 f
t
)]dk
t
,
where f = f(x, k, t), f
t
= f(x, k
t
, t). Equation (3.23), together with the eective eld
equation
E
e
(x, t) = E
ext
(x, t) +
_
R
d
n(y, t)E
int
(x, y) dy, (3.25)
where E
ext
and E
int
are given functions, and the collision operator (3.24), is called the
semi-classical Boltzmann equation. When E
ext
and E
int
are given by the Coulomb forces
(3.19) and (3.20), equations (3.23)(3.25) are called the Boltzmann-Poisson system which
can be written as (3.21)(3.22) with f instead of F and with the right-hand side Q(f) in
(3.21).
Again we impose the initial and periodic boundary conditions
f(x, k, t) = f(x, k, t), x R
d
, k B, t > 0, (3.26)
f(x, k, 0) = f
I
(x, k), x R
d
, k B. (3.27)
The Boltzmann equation has two nonlinearities:
a quadratic nonlocal nonlinearity in the position variable caused by the self-consistent
eld E
e
in (3.25) and
another quadratic nonlocal nonlineariy in the wave vector caused by the collision
integral (3.24).
These nonlinearities make the mathematical analysis of the initial-boundary-value prob-
lem (3.23)(3.27) very dicult. Only at the end of the 1980ies, an existence proof for
global-in-time solutions and large data in the eld-free case E
e
= 0 has been obtained
by DiPerna and Lions employing the theory of renormalized solutions [26, 27]. We refer
to the reviews [44, 63] for details and more references.
We give now some examples of collision operators.
Example 3.6 In semiconductor crystals, scattering of electrons occurs by lattice defects,
phonons, and other carriers. We consider only the following important collision events:
electron-phonon scattering,
ionized impurity scattering, and
59
carrier-carrier scattering.
Extensive treatments of scattering mechanisms in semiconductors can be found in, e.g.,
[11, Ch. 9], [37, Ch. 7], [45, Ch. 2], [61, Ch. 6], and in the textbooks [33, 58, 68].
Phonon scattering. At nonzero temperature, the atoms in the crystal lattice vibrate
about their xed equilibrium. These vibrations are quantized and the quantum of lattice
vibrations is called phonon. We can distinguish so-called acoustic phonons and optical
phonons. Acoustic phonons arise from displacements of lattice atoms in the same direction
like sound waves. Optical phonos describe displacements in the wave vector and are able
to interact strongly with light. Denoting by

the energy of a phonon, the phonon


occupation number N

is given by Bose-Einstein statistics,


N

=
1
e
/k
B
T
1
,
where the index refers to either op for optical phonons or ac for accoustic phonons.
Notice that Bose-Einstein statistics can be used for indistinguishable particles not obey-
ing the Pauli exclusion principle and therefore also for phonons (see [11, p.307.] for a
derivation).
An electron in the Bloch state k
t
with conduction-band energy E(k
t
) before the colli-
sion with a phonon can change to the state k after the collision if
E(k
t
) E(k) =

, (3.28)
where the plus sign refers to the phonon emission and the minus sign for phonon absorp-
tion. The acoustic phonon energy

is usually a function of the dierence k k


t
of the
wave vectors before and after a scattering event [8, Sec. 2]. The transition rate s(x, k, k
t
)
is non-zero only if (3.28) is satised. Therefore,
s

(x, k, k
t
) =

(x, k, k
t
)[(1 +N

)(E(k
t
) E(k) +

) + N

(E(k
t
) E(k) )],
where

(x, k, k
t
) is assumed to be symmetric in k and k
t
and is the delta distribution
used in Section 2.4. This transition rate can be derived more generally from Fermis golden
rule (see [11, Sec. 4.4], [45, Sec. 1.7.1], or [65, Appendix C]). The collision operator reads
according to (3.24)
(Q

(f))(x, k, t) =
_
B
[s

(x, k
t
, k)f
t
(1 f) s

(x, k, k
t
)f(1 f
t
)] dk
t
. (3.29)
The above expression for s

shows that the scattering rates can be highly non-smooth.


Ionized impurity scattering. A doping atom in the semiconductor material donates
either an electron or a hole, leaving behind an ionized charged impurity center. This xed
charge may attract or repulse an electron propagating through the crystal lattice. The
60
interaction of carriers with neutral impurities is another scattering possibility but we do
not consider it here. Since the scattering is elastic, the electron energy E(k
t
) after the
collision is the same as the energy E(k) before the interaction, and the transition rate is
[57]
s
imp
(x, k, k
t
) =
imp
(x, k, k
t
)(E(k
t
) E(k)),
where
imp
is symmetric in k and k
t
. Again, this expression can be derived from Fermis
golden rule (see above). The symmetry of
imp
and implies that
s
imp
(x, k
t
, k)f
t
f s
imp
(x, k, k
t
)ff
t
=
imp
(x, k, k
t
)(E(k
t
) E(k))(f
t
f ff
t
) = 0,
and hence, the collision operator becomes
(Q
imp
(f))(x, k, t) =
_
B

imp
(x, k, k
t
)(E(k
t
) E(k))(f
t
f) dk
t
.
Carrier-carrier scattering. We only consider binary electron-electron interactions. Also
binary electron-hole scattering or collective carrier-carrier collisions (i.e. interactions of
carriers with oscillations in the carrier density; see [45, Sec. 2.10.2]) are possible but we do
not consider these mechanisms here. The inuence of electron-electron interactions on the
carrier dynamics is more pronounced in degenerate semiconductors in which Fermi-Dirac
statistics instead of Maxwell-Boltzmann statistics has to be used (see Section 2.4). The
transition rate that carrier in the Bloch states k
t
and k
t
1
collide and scatter to the states
k and k
1
is given by
s
ee
(x, k, k
t
, k
1
, k
t
1
) =
ee
(x, k, k
t
, k
1
, k
t
1
)(E(k
t
) + E(k
t
1
) E(k) E(k
1
)),
since the collisions are elastic. Therefore, the collision operator becomes [9, (2.7)]:
(Q
ee
(f))(x, k, t) =
_
B
3
s
ee
(x, k, k
t
, k
1
, k
t
1
)[f
t
f
t
1
(1f)(1f
t
)ff
1
(1f
t
)(1f
t
1
)]dk
t
dk
1
dk
t
1
,
(3.30)
where f = f(x, k, t), f
t
= f(x, k
t
, t), f
1
= f(x, k
1
, t), and f
t
1
= f(x, k
t
1
, t). Notice that this
operator has a nonlocal nonlinearity of fourth order.
Generally, the collision operator in (3.23) can be written as the sum of the various
collision operators considered above:
Q(f) = Q
op
(f) + Q
ac
(f) + Q
imp
(f) + Q
ee
(f).
In the following we show some properties of the collision operator (3.24). The transition
rate s(x, k, k
t
) is assumed to satisfy
s(x, k, k
t
)
s(x, k
t
, k)
= exp
E(k) E(k
t
)
k
B
T
for x R
d
, k, k
t
B. (3.31)
61
We explain this hypothesis. The so-called principle of detailed balance [3, 37] asserts that
in thermal equilibrium, the local scattering probabilities vanish,
s(x, k
t
, k)f
t
eq
(1 f
eq
) s(x, k, k
t
)f
eq
(1 f
t
eq
) = 0,
and that the equilibrium occupation number density is given by the Fermi-Dirac distri-
bution,
f
eq
(x, k) =
1
1 + e
u(k)
, u(k) =
E(k) E
F
k
B
T
(see (2.48)). A calculation shows that
s(x, k, k
t
)
s(x, k
t
, k)
=
f
t
eq
(1 f
eq
)
f
eq
(1 f
t
eq
)
= e
u(k)u(k

)
,
which is (3.31).
The following result is due to Poupand [52] (also see the review [54]).
Theorem 3.7 Let (3.31) hold for some function E(k) and let s(x, k, k
t
) > 0 for all
x R
d
, k, k
t
B.
(1) For all (regular) functions f there holds:
_
B
(Q(f))(x, k, t) dk = 0 for x R
d
, t > 0.
(2) For all functions f(x, k, t) (0, 1) and non-decreasing functions : R R there
holds:
_
B
(Q(f))(x, k, t)
_
f(x, k, t)
1 f(x, k, t)
e
E(k)/k
B
T
_
dk 0, (3.32)
_
B
(Q(f))(x, k, t)
_
1 f(x, k, t)
f(x, k, t)
e
E(k)/k
B
T
_
dk 0.
(3) The kernel of Q only consists of Fermi-Dirac distributions, i.e. Q(f) = 0 if and only
if, for some E
F
,
f(k) =
1
1 + exp((E(k) E
F
)/k
B
T)
, k B. (3.33)
The theorem can be physically interpreted as follows. Let f be a solution to the
Boltzmann equation (3.23)(3.24). Then, by Theorem 3.7 (1),

t
_
R
d
n(x, t) dx =
_
R
d
_
B

t
f(x, k, t) dk dx
=
_
R
d
_
B
(div
x
(v(k)f) +
q
h
div
k
(E
e
f) + Q(f)) dk dx
= 0,
62
where we have used the divergence theorem. This implies that the total number of elec-
trons is conserved in time:
_
R
d
n(x, t) dx =
_
R
d
n(x, 0) dx for all t > 0.
Physically, this is reasonable: collisions neither destroy nor generate particles. The state-
ment of Theorem 3.7 (2) is also called H-theorem.
Proof of Theorem 3.7: (1) Changing k and k
t
in the second integral gives
_
B
Q(f) dk =
_
B
2
s(k
t
, k)f
t
(1 f) dk
t
dk
_
B
2
s(k, k
t
)f(1 f
t
) dk
t
dk
=
_
B
2
s(k
t
, k)f
t
(1 f) dk
t
dk
_
B
2
s(k
t
, k)f
t
(1 f) dk dk
t
= 0.
(2) We show only the second inequality. The proof of the rst one is similar. Set
M(k) = e
E(k)/k
B
T
, F(k) =
1 f(k)
f(k)
M(k). (3.34)
The function M(k) is called Maxwellian. Then assumption (3.31) is equivalent to
s(k
t
, k)
M(k)
=
s(k, k
t
)
M(k
t
)
, (3.35)
and we obtain, with the notations F = F(k), F
t
= F(k
t
), M = M(k), M
t
= M(k
t
),
_
B
Q(f)(F) dk =
_
B
2
s(k, k
t
)
_
M
M
t
f
t
(1 f)(F) f(1 f
t
)(F)
_
dk
t
dk
=
_
B
2
s(k, k
t
)
M
t
ff
t
(F F
t
)(F) dk
t
dk, (3.36)
and, after changing k and k
t
and then using (3.35),
_
B
Q(f)(F) dk =
_
B
2
s(k
t
, k)
M
f
t
f(F
t
F)(F
t
) dk dk
t
=
_
B
2
s(k, k
t
)
M
t
f
t
f(F
t
F)(F
t
) dk
t
dk. (3.37)
Adding (3.36) and (3.37) leads to
_
B
Q(f)(F) dk =
1
2
_
B
2
s(k, k
t
)
M
t
ff
t
(F F
t
)((F) (F
t
)) dk
t
dk
0, (3.38)
since is non-decreasing and all the other factors are non-negative.
63
(3) We see from (2), with (x) = x, taking into account (3.38), that Q(f) = 0 is
equivalent to
ff
t
(F F
t
)
2
= 0 for almost all k, k
t
B.
This implies f 0 or F = F
t
almost everywhere. The latter equation is equivalent to
1 f(k)
f(k)
M(k) =
1 f(k
t
)
f(k
t
)
M(k
t
) for almost all k, k
t
B.
We infer that both sides are constant and denote this constant by exp(E
F
/k
B
T) with
E
F
R. Notice that the constant is positive since 0 f 1, except if f 1. But then
1 f(k)
f(k)
=
e
E
F
/k
B
T
M(k)
= e
(E(k)E
F
)/k
B
T
and solving for f(k) yields (3.33). Finally, choosing E
F
= leads to the other two
possibilities F 0 of F 1.
In the literature, two approximations of the collision operator are frequently used [49,
p. 33]:
the low density approximation
Q
0
(f))(x, k, t) =
_
B
(x, k
t
, k)(Mf
t
M
t
f) dk
t
, (3.39)
where (x, k
t
, k) = s(x, k
t
, k)/M(k) is called collision cross-section;
the relaxation time approximation
(Q

(f))(x, k, t) =
1
(x, k)
(f(x, k, t) M(k)n(x)) , (3.40)
where
(x, k) =
__
B
s(x, k, k
t
) dk
t
_
1
is called the relaxation time describing the average time between two consecutive
scattering events at (x, k), and n(x) is some given density such that
n(x) =
_
B
f
I
(x, k) dk.
These approximations can be derived as follows. The low-density collision operator is
obtained by assuming that the distribution function is small:
0 f(x, k, t) 1.
64
Then 1 f(x) 1 and we can approximate, using (3.35),
Q(f)(k) =
_
B
[s(k
t
, k)f
t
s(k, k
t
)f] dk
t
=
_
B
s(k
t
, k)
M(k)
[M(k)f
t
M(k
t
)f] dk
t
,
which is (3.39). Notice that (3.35) implies that (x, k
t
, k) is symmetric in k and k
t
.
When the initial distribution is close to a multiple of M(k) it is reasonable to approx-
imate f
t
in (3.39) by n(x)M(k
t
). Thus, (3.39) yields
(Q(f))(k) =
_
B
(k
t
, k)(nMM
t
M
t
f) dk
t
=
_
B
s(k
t
, k)
M
t
M
(nM f) dk
t
=
_
B
s(k, k
t
)(nM f) dk
t
= (nM f)
_
B
s(k, k
t
) dk
t
,
which is equal to (3.40).
Why is (x, k) called relaxation time? In order to see this, we argue as in [49, p. 34f.]
and consider the Boltzmann equation with (3.40) along the trajectories (x(t), k(t))
df
dt
=
1

(f nM), t > 0,
where x(t), k(t) solves x = v(k),

k = qE
e
/, x(0) = x
0
, k(0) = k
0
for constant relax-
ation time . Multiplying this dierential equation with f nM gives
1
2
d
dt
(f nM)
2
= (f nM)
df
dt
=
1

(f nM)
2
.
Separation of variables leads to
f(x(t), k(t), t) n(x(t))M(k(t)) = ce
t/
0 as t ,
where c R is an integration constant. This means that the distribution function relaxes
to the equilibrium density nM, starting from f
I
, along the trajectory after a time of order
.
Notice that the low-density and the relaxation-time approximations (3.39) and (3.40)
coincide if the collision cross-section only depends on the position variable and if the
Maxwellians are normalized such that
_
B
M dk = 1. Indeed, we obtain
(Q
0
(f))(x, k, t) =
_
B
(x)(Mf
t
M
t
f) dk
t
= (x)M(k)
_
B
f
t
dk
t
(x)f(x)
_
B
M
t
dk
t
= (x)(f(x, k, t) M(k)n(x, t)),
65
which corresponds to (3.40) with (x) = 1/(k) and time-dependent density n(x, t).
Lemma 3.8 Let (x, k
t
, k) in (3.39) be symmetric in k and k
t
. Then the kernel of the
low-density collision operator consists exactly of the Maxwellians, i.e. Q
0
(f) = 0 if and
only if
f(x, k) = n(x)M(k) = n(x)e
E(k)/k
B
T
,
where n(x) is any (non-negative) function.
Proof: The proof is very similar to the proof of Theorem 3.7, parts (2) and (3), using
(x) = x and F(k) = M(k)/f(k). Indeed, an analogous computation leads to F(k) =
F(k
t
) or
M(k)
f(k)
=
M(k
t
)
f(k
t
)
for almost all k, k
t
B.
This implies M(k)/f(k) = const. = 1/n(x).
Remark 3.9 In the parabolic band approximation
E(k) =

2
2m

[k k
0
[
2
we can characterize the kernel of the low-density operator by the family of functions
M
,u,T
(v) =
_
m

2k
B
T
_
d/2
exp
_

[v u[
2
2k
B
T
_
(3.41)
with , T > 0 and u R. Indeed, the velocity is
1

k
E(k) =

m

(k k
0
) =: v u
such that
M(v) = exp
_

E(k)
k
B
T
_
= exp
_

[v u[
2
2k
B
T
_
.
Setting
n(x) =
_
m

2k
B
T
_
d/2
then gives (3.41). The constant represents a density, u a velocity, and T the temperature.

The semi-classical Boltzmann equation (3.23) is fundamental in deriving simpler macro-


scopic models for semiconductors (see Chapter 3). It is the basis equation in (classical)
semiconductor modeling and usually, other models are validated by numerical comparisons
with the Boltzmann equation. Nevertheless, it is important to understand its limitations
(see [45, Sec. 37]):
66
The Boltzmann equation is only a single-particle description of many-particle sys-
tems of charged carriers. In particular, correlations between carriers are neglected.
Quantum mechanical phenomena are only modeled in a semi-classical way. Electrons
are considered as particles which obey Newtons laws.
Collisions are assumed to be binary and to be instantaneous in time and local in
space.
We can estimate the range of validity by using Heisenbergs uncertainty principles
px and Et , (3.42)
which means that it is impossible to determine both momentum and space at the same
time and that the energy of a particle can be dened only if it stays in the same state for
some time [45, p. 152]. For instance, writing
p
2
2m

= E
for the momentum p and assuming that the energy spread is of the order of the thermal
energy k
B
T, we obtain
p =
_
2m

E
_
2m

k
B
T.
Introduce the de Broglie length

B
=
h

2m

k
B
T
which is the wave length of an electron with thermal energy. Then the rst inequality in
(3.42) leads to the requirement
x

p


B
2
.
Therefore, when treating electrons as particles, they cannot be localized sharper than x
which is of the order of
B
. At room temperature and with the eective mass m

= 0.98m
e
(see Table 2.3) of silicon, we obtain x 8 nm.
Taking t to be the time between collisions and again assuming that E k
B
T, the
second inequality in (3.42) gives
t

E


k
B
T
.
Supposing further that the electron has the velocity v = p/m

_
2k
B
T/m

, correspond-
ing to the thermal energy, the distance between two collisions is
L = vt
_
2k
B
T
m

2k
B
T
=

B

.
67
Thus, the mean free path L should be (much) larger than the de Broglie length
B

8 nm.
Finally, in Figure 3.2 we present a summary of the models derived in this chapter and
the relations between them.
semi-classical Boltzmann equation
semi-classical Vlasov equation
semi-classical Liouville equation
no collisions
no two-particle interactions
Figure 3.2: Relations between the models of Sections 3.13.3.
3.4 The bipolar Boltzmann model
So far we have only considered the transport of electrons in the conduction band. How-
ever, also holes in the valence band contribute to the carrier ow in semiconductors (see
Section 2.2). It is possible that an electron moves from the valence band to the conduction
band, leaving a hole in the valence band behind it. This process is called the generation
of an electron-hole pair (see Figure 3.3). The electron has to overcome the energy gap,
which is of the order of 1eV; on the other hand, the thermal energy of an electron is
only of the order of k
B
T 0.026eV at room temperature. Therefore, a lot of absorption
energy is necessary for such processes. The inverse process of an electron moving from the
conduction to the valence band, occupying an empty state, is termed the recombination
of an electron-hole pair. For such an event, energy is emitted. The basic mechanisms for
generation-recombination processes are
Auger/impact ionization generation-recombination,
radiative generation-recombination,
thermal generation-recombination.
68
An Auger process is dened as an electron-hole recombination followed by a transfer of
energy to a free carrier which is excited to a state of higher energy. The inverse Auger
process, i.e. the generation of an electron-hole pair, is called impact ionization. The energy
for the pair generation comes from the collision of a high-energetic free carrier with the
lattice or from electron-electron or hole-hole collisions.
When an electron from the conduction band recombines with a hole from the valence
band and emittes a photon, we call this process radiative recombination. The energy of
the photon is equal to the bandgap energy. Radiative generation occurs when a photon
with energy larger than or equal to the gap energy is absorbed.
A third source of energy is coming from lattice vibrations or phonons. Thus, thermal
recombination or generation arises from phonon emission or absorption, respectively.

` E(k)
k

valence band
conduction band
energy
emission

` E(k)
k
`
valence band
conduction band
energy
absorption

Figure 3.3: Recombination (left) and generation (right) of an electron-hole pair.


The recombination-generation operators can be derived as the collision operator of
Section 3.3 from phenomenological considerations. The generation of an electron in the
state k and a hole in the state k
t
is possible if both states are not occupied, and its rate
is given by
g(x, k
t
, k)(1 f
n
)(1 f
t
p
),
where g(x, k
t
, k) 0 is the generation rate, f
n
= f
n
(x, k, t) the electron distribution
function and f
t
p
= f
p
(x, k
t
, t) the hole distribution function. The rate of recombination of
an electron at state k and a hole at state k
t
is
r(x, k, k
t
)f
n
f
t
p
,
where r(x, k, k
t
) 0. The net rate is the dierence of generation and recombination rates:
g(x, k
t
, k)(1 f
n
)(1 f
t
p
) r(x, k, k
t
)f
n
f
t
p
,
69
and the recombination-generation operator for electrons in the conduction band is the
integral over all states k
t
(see [49, Sec. 1.6]):
(I
n
(f
n
, f
p
))(x, k, t) =
_
B
[g(x, k
t
, k)(1 f
n
)(1 f
t
p
) r(x, k, k
t
)f
n
f
t
p
] dk
t
. (3.43)
In a similar way, the recombination-generation operator for holes in the valence band can
be written as
(I
p
(f
n
, f
p
))(x, k, t) =
_
B
[g(x, k, k
t
)(1 f
t
n
)(1 f
p
) r(x, k
t
, k)f
t
n
f
p
] dk
t
. (3.44)
The recombination and generation rates are related by the equation
r(x, k, k
t
) = exp
_
E
n
(k) E
p
(k
t
)
k
B
T
_
g(x, k
t
, k) (3.45)
which can be derived from the principle of detailed balance as in Section 3.3 (see (3.39)).
This principle holds here since recombination and generation balance in thermal equilib-
rium.
The operators (3.43) and (3.44) are added to the electron and hole collision operators.
Then, the evolution of the distribution functions f
n
and f
p
is given by the system of
Boltzmann equations

t
f
n
+ v
n
(k)
x
f
n

E
e

k
f
n
= Q
n
(f
n
) + I
n
(f
n
, f
p
) (3.46)

t
f
p
+ v
p
(k)
x
f
p
+
q

E
e

k
f
p
= Q
p
(f
p
) + I
p
(f
n
, f
p
), (3.47)
where
v
n
(k) =
1

k
E
n
(k), v
p
(k) =
1

k
E
p
(k)
and E
n
and E
p
are the conduction and valance band energies, respectively. Denoting by
n(x, t) =
_
B
f
n
(x, k, t) dk, p(x, t) =
_
B
f
p
(x, k, t) dk
the electron and hole densities, respectively, the eective-eld equation (3.25) becomes
E
e
(x, t) = E
ext
(x, t) +
_
R
d
(n(y, t) p(y, t))E
int
(x, y) dy, (3.48)
since electrons and holes have charges with opposite sign. The equations (3.46)(3.48),
together with the collision operators (3.24) and (3.43)(3.44) are called the semi-classical
bipolar Boltzmann model.
The bipolar model has an additional nonlinearity due to the coupling between the
carrier densities n and p through (3.48). In particular, the total number of each type of
particles is not conserved anymore since recombination-generation eects can take place.
70
However, the total space charge np C is conserved if the doping atoms are immobile,
i.e., C is a function of the space variable x only. Indeed, taking the dierence of the
Boltzmann equations (3.46) and (3.47) and integrating over (x, k) R
d
B yields

t
_
R
d
(n p C(x)) dx =
_
R
d
_
B
(I
n
(f
n
, f
p
) I
p
(f
n
, f
P
)) dk dx = 0,
by arguing as in the proof of Theorem 3.7 (1).
Finally, we discuss two special cases. In the low density approximation f
n
, f
p
1 we
can write the recombination-generation operators as
(I
n
(f
n
, f
p
))(x, k, t) =
_
B
g(x, k
t
, k)
_
1 e
(En(k)Ep(k

))/k
B
T
f
n
f
t
p
_
dk
t
, (3.49)
(I
p
(f
n
, f
p
))(x, k, t) =
_
B
g(x, k, k
t
)
_
1 e
(En(k

)Ep(k))/k
B
T
f
t
n
f
p
_
dk
t
, (3.50)
using (3.45). In the case of Coulomb forces in R
3
, the eective eld is given by
E
e
(x, t) =
1
4
s
_
R
3
(y, t)
x y
[x y[
3
dy,
where the total space charge is the sum of the electron density n, the hole density
p, and the densities N
D
, N
A
of the implanted positively charged donor ions and the
negatively charged acceptor ions, respectively, with which the semiconductor is doped
(see Section 2.4), weighted by the corresponding charges +q or q:
= q(n + p N
A
+ N
D
).
Dening the electrostatic potential E
e
= V and the doping prole C = N
D
N
A
, we
can replace the eective-eld equation (3.48) by the Poisson equation

s
V = q(n p C) in R
3
. (3.51)
The Boltzmann equations (3.43)(3.44) and the Poisson equation (3.51) constitute the
so-called bipolar Boltzmann-Poisson system.
71
4 Classical Fluid Models
The mathematical and numerical solution of the Boltzmann equation of Chapter 3 is
very dicult due to the complex structure of the collision operator and the high number
of independent variables (three position plus three wave vector plus one time variable).
In this chapter we derive simpler models with only four independent variables (three
space and one time variable) from the Boltzmann equation. The main feature of these
models is that they describe the evolution of averaged quantities, like the carrier density
or the current density, rather than a distribution function. Thus the carrier ensemble is
considered as a uid consisting of charged particles.
4.1 Derivation of the drift-diusion equations
The drift-diusion equations are one of the simplest semiconductor models. It has been
rst derived by van Roosbroeck in 1950 [64]. We derive this model by employing the
so-called Hilbert method which consists in expanding the distribution function in powers
of the scaled mean free path. This procedure has been analyzed in [51]. We follow the
presentation of [49, Sec. 22].
The starting point of the derivation is the bipolar Boltzmann model of Section 3.4 in
three dimensions,

t
f
n
+ v
n
(k)
x
f
n

E
e

k
f
n
= Q
n
(f
n
) + I
n
(f
n
, f
p
), (4.1)

t
f
p
+ v
p
(k)
x
f
p
+
q

E
e

k
f
p
= Q
p
(f
p
) + I
p
(f
n
, f
p
), (4.2)
together with the eld equation

s
divE
e
= q(n p C(x)), x, k R
3
.
We assume low-density collision operators,
(Q
j
(f
j
))(x, k, t) =
_
B

j
(x, k, k
t
)
_
e
E
j
(k)/k
B
T
f
t
j
e
E
j
(k

)/k
B
T
f
j
_
dk
t
, (4.3)
where j n, p (see (3.39)) and the low-density recombination-generation operators
(3.49)(3.50),
(I
n
(f
n
, f
p
))(x, k, t) =
_
B
g(x, k
t
, k)
_
1 e
(En(k)Ep(k

))/k
B
T
_
f
n
f
t
p
dk
t
, (4.4)
(I
p
(f
n
, f
p
))(x, k, t) =
_
B
g(x, k, k
t
)
_
1 e
(En(k

)Ep(k))/k
B
T
_
f
n
f
t
p
dk
t
. (4.5)
As in Chapter 3, f
t
j
means evaluation at k
t
, i.e., f
t
j
= f
j
(x, k
t
, t), j n, p, and B R
3
denotes the Brillouin zone. Furthermore, we assume the parabolic band approximation
E
n
(k) = E
c
+

2
2m

e
[k[
2
, E
p
(k) = E
v


2
2m

h
[k[
2
,
72
where E
c
denotes the conduction band minimum, E
v
the valence band maximum, and m

e
and m

h
are the eective masses of electrons and holes, respectively. In particular, we can
replace B by R
3
in the above integrals. The velocities in (4.1)(4.2) are given by
v
n
(k) =
1

k
E
n
(k) =
k
m

e
, v
p
(k) =
1

k
E
p
(k) =
k
m

h
.
The drift-diusion equations are derived under the (main) assumption that collisions
occur on a much shorter time scale than recombination-generation events. Indeed, typ-
ically the time scale of collisions is 10
12
s whereas a typical time for recombination-
generation eects is 10
9
s [49, p. 86]. In order to make this statement more precise, we
scale the Boltzmann equations, i.e., we choose new variables which make the equation
dimension free. We choose the following reference parameter:
Dene the reference velocity v =
_
k
B
T/m

e
. Behind this setting there are two
assumptions: The eective masses of electrons and holes are of the same order and
the thermal energy k
B
T is of the same order as the kinetic energy m

e
v
2
/2.
Let
c
,
R
be characteristic times of collision and recombination-generation events.
Then
c
=
c
v and
R
=
R
v are the so-called mean free paths between two consec-
utive scattering and recombination-generation events. Thus we have to scale
Q
n
=
1

c
Q
n,s
, Q
p
=
1

c
Q
p,s
,
I
n
=
1

R
I
n,s
, I
p
=
1

R
I
p,s
.
A reference length
0
is given by the geometric average of
c
and
R
:

0
=
_

c
.
Then x =
0
x
s
.
We use the reference time
R
, the reference wave vector m

e
v/ and the reference
eld strength k
B
T/q
0
:
t =
R
t
s
, k =
m

e
v

k
s
, E
e
=
k
B
T

0
q
E
e,s
. (4.6)
Our main assumption is

c

R
.
Then the parameter
2
=
c
/
R
=
c
/
R
satises 1. With the above scaling we can
rewrite (4.1) as
1

ts
f
n
+
v

0
k
s

xs
f
n

k
B
T

0
m

e
v
E
e,s

ks
f
n
=
1

c
Q
n,s
(f
n
) +
1

R
I
n,s
(f
n
, f
p
).
73
Multiplying this equation by
c
=
c
/v, using =
c
/
0
and k
B
T = m

e
v
2
, and omitting
the index s, we obtain the scaled Boltzmann equation

t
f
n
+ (k
x
f
n
E
e

k
f
n
) = Q
n
(f
n
) +
2
I
n
(f
n
, f
p
). (4.7)
In a similar way, we have

t
f
p
+
_
m

e
m

h
k
x
f
p
+ E
e

k
f
p
_
= Q
p
(f
p
) +
2
I
p
(f
n
, f
p
). (4.8)
The scaled operators Q
j,s
and I
j,s
have the same form as (4.3)(4.5) but
e
En(k)/k
B
T
and e
Ep(k)/k
B
T
have to be replaced by
e
Ec/k
B
T[k[
2
/2
and e
Ev/k
B
T(m

e
/m

h
)[k[
2
/2
,
respectively, and the rates
n
,
p
, and g are multiplied by (m

e
v/)
d
.
We want to study the scaled Boltzmann equations for small . First we analyze
the collision operators Q
n
and Q
p
. We proceed as in [53]. It is convenient to write them
in the form
Q
f
(f
j
) =
_
R
3
s
j
(x, k, k
t
)(M
j
(k)f
t
j
M
j
(k
t
)f
j
) dk
t
, j n, p, (4.9)
where
s
n
(x, k, k
t
) = N
n

n
(x, k, k
t
)e
Ec/k
B
T
,
s
p
(x, k, k
t
) = N
p

p
(x, k, k
t
)e
Ev/k
B
T
,
and
M
n
(k) =
1
N
n
e
[k[
2
/2
, M
p
(k) =
1
N
n
e
(m

e
/m

h
)[k[
2
/2
are the scaled Maxwellians. The constants
N
n
= (2)
3/2
, N
p
= (2)
3/2
_
m

h
m

e
_
3/2
are chosen in such a way that the Maxwellians are normalized, i.e.
_
R
3
M
n
(k) dk =
_
R
3
M
p
(k) dk = 1.
Here we have used that fact that
_
R
d
e
[k[
2
/2
dk = (2)
d/2
.
74
For the following analysis we introduce the functions

j
(k) =
_
R
3
s
j
(x, k, k
t
)M
j
(k
t
) dk
t
, k R
3
, j n, p, (4.10)
for some xed x R
d
, and the Banach spaces
X
j
= f : R
3
R measurable: |f|
X
j
< ,
Y
j
= f : R
3
R measurable: |f|
Y
j
<
with associated norms
|f|
X
j
=
__
R
3
f(k)
2

j
(k)M
j
(k)
1
dk
_
1/2
,
|f|
Y
j
=
__
R
3
f(k)
2

j
(k)
1
M
j
(k)
1
dk
_
1/2
.
Lemma 4.1 Let j n, p and let s
j
> 0 be symmetric in k and k
t
. Then the kernel of
Q
j
, N(Q
j
) = f X
j
: Q
j
(f) = 0, only consists of Maxwellians:
N(Q
j
) = M
j
: = (x) R.
Proof: The proof is similar to the proofs of Theorem 3.7 and Lemma 3.8. In the following
we omit the index j. First we symmetrize the collision operator by setting f
s
= (/M)
1/2
f
and Q
s
(f
s
) = (M)
1/2
Q(f). Then
Q
s
(f
s
) = (M)
1/2
_
M
_
R
3
s(x, k, k
t
)f
t
dk
t
f
_
=
_
R
3
s(x, k, k
t
)
_
MM
t

t
_
1/2
f
t
s
dk
t
f
s
,
where M
t
= M(k
t
) and
t
= (k
t
). Since s is symmetric in k and k
t
by assumption, the
operator Q
s
: L
2
(R
3
) L
2
(R
3
) is self-adjoint (see Section 2.1).
Now we analyze the kernel of Q
s
. By denition of we have
_
R
3
s(x, k, k
t
)
M
t

dk
t
= 1
75
and therefore

_
R
3
Q
s
(f
s
)f
s
dk =
_
R
3
_
f
2
s

_
R
3
s(x, k, k
t
)
_
MM
t

t
_
1/2
f
t
s
f
s
dk
t
_
dk
=
1
2
_
R
3
__
R
3
s(x, k, k
t
)
M
t

dk
t
_
f
2
s
dk
+
1
2
_
R
3
__
R
3
s(x, k, k
t
)
M

t
dk
_
(f
t
s
)
2
dk
t

_
R
3
_
R
3
s(x, k, k
t
)
_
MM
t

t
_
1/2
f
t
s
f
s
dk
t
dk
=
1
2
_
R
3
_
R
3
s(x, k, k
t
)MM
t
_
f
s

f
t
s

t
M
t
_
2
dk dk
t
0. (4.11)
Thus, Q
s
(f
s
) = 0 implies that
f
s

M
=
f
t
s

t
M
t
for k, k
t
R
3
.
We conclude that both sides must be constant, f
s
/

M = = const., where = (x)


is a parameter. In the original variables, Q(f) = 0 implies f =
_
M/f
s
=
_
M/

M = M for all R. Conversely, if f = M for some R, the formulation (4.9)


immediately gives Q(f) = 0. This proves the lemma.
Lemma 4.2 Let the assumptions of Lemma 4.1 hold. Then
(1) The equation Q
j
(f) = g Y
j
has a solution f X
j
if and only if
_
R
3
g(k) dk = 0.
In this situation, any solution of Q
j
(f) = g can be written as f + M
j
, where
= (x) is a parameter.
(2) The solution f X
j
of Q
j
(f) = g is unique if the orthonogality relation
_
R
3
f(k)
j
(k) dk = 0 (4.12)
is satised.
This lemma is a consequence of the Fredholm alternative which we prove rst.
76
Lemma 4.3 (Fredholm alternative)
Let X be a Hilbert space with scalar product (, ) and Q : X X a linear, continuous
and closed operator (i.e. R(Q) = g X : f X : Q(f) = g is closed). Then
Q(f) = g has a solution g N(Q

.
Proof: We show rst that N(Q

) = R(Q)

. Indeed,
f N(Q

) Q

(f) = 0 (g, Q

(f)) = 0 for all g X


(Q(g), f) = 0 for all g X f R(Q)
f R(Q)

.
Thus, since Q is closed,
R(Q) = R(Q) = R(Q)

= N(Q

.
This identity is equivalent to the assertion.
Proof of Lemma 4.2: Again, we omit the index j.
(1) It is possible to show that the operator Q
s
: L
2
(R
3
) L
2
(R
3
) is linear, continuous,
and closed. Moreover, we have shown in the proof of Lemma 4.1 that Q
s
is self-adjoint. We
conclude from Lemma 4.3 that Q
s
(f
s
) = g
s
has a solution if and only if g
s
N(Q

s
)

=
N(Q
s
)

or
_
R
3
g
s
hdk = 0 for all h N(Q
s
).
By Lemma 4.1, the kernel of Q
s
is spanned by

M, such that this equivalent to


0 =
_
R
3
g
s

M dk =
_
R
d
g dk,
since in the original variables we have
g = Q(f) =

MQ
s
(f
s
) =

Mg
s
.
Let f
1
and f
2
be two solutions of Q(f) = g. Then, since Q is linear, Q(f
1
f
2
) =
Q(f
1
) Q(f
2
) = 0 and f
1
f
2
N(Q). This shows that f
1
= f
2
+M for some R.
(2) We only give a sketch of the proof and refer to [53] for details. It is possible to
show that the operator Q is coercive in the following sense:
_
R
3
(Q(f))fM
1
dk c|f|
2
X
(4.13)
for some c > 0 and for all f X satisfying (4.12). Clearly, this implies that Q is one-to-
one on the subset of functions satisfying (4.12), and the uniqueness property is shown. In
order to prove (4.13) one needs to show that Id +Q is a Hilbert-Schmidt operator and to
use general properties of these operators (see [53]).
77
Lemma 4.4 Let the assumptions of Lemma 4.1 hold and assume that
s
j
(x, Ak, Ak
t
) = s
j
(x, k, k
t
) for x, k, k
t
R, j n, p, (4.14)
and for all isometric matrices A R
33
. Then the equations
(Q
n
(h
n,i
))(k) = k
i
M
n
(k), (Q
p
(h
p,i
))(k) =
m

e
m

h
k
i
M
p
(k), i = 1, 2, 3,
have solutions h
n
(x, k) = (h
n,1
(x, k), . . . , h
n,3
(x, k)) and h
p
(x, k) = (h
p,1
(x, k), . . . , h
p,3
(x, k))
with the property that there exist
n
(x),
p
(x) 0 satisfying
_
R
3
k h
n
dk =
n
(x)I, (4.15)
_
R
3
m

e
m

h
k h
p
dk =
p
(x)I, (4.16)
where I R
33
is the unit matrix and a b = a

b R
33
for a, b R
3
.
In the statement of the lemma we have omitted some technical assumptions on the
scattering rates s
n
and s
p
(regularity conditions). Again, we refer to [53] for the precise
hypotheses.
Proof: Again we omit the index j n, p. The existence of a solution h(k) = (h
1
(k),
h
2
(k), h
3
(k)) of Q(h) = kM(k) follows from Lemma 4.3 (1) since
_
R
3
k
i
M(k) dk = 0 for all i = 1, 2, 3.
We only show (4.15) since the proof of (4.16) is similar.
Let A be the matrix of a rotation with axis k
1
. Then (Ak)
1
= k
1
for all k =
(k
1
, k
2
, k
3
)

R
3
. Since A is isometric, i.e., [Ak[ = [k[, we obtain M(Ak) = N
1
e
[Ak[
2
/2
=
N
1
e
[k[
2
/2
= M(k) and (Ak)
1
M(Ak) = k
1
M(k). This implies, together with the assump-
tion (4.14) and the transformation w = Ak
t
with dw = [ det A[dk
t
= dk
t
(Q(h
1
A))(k) =
_
R
3
s(x, k, k
t
)[M(k)h
1
(Ak
t
) M(k
t
)h
1
(Ak)] dk
t
=
_
R
3
s(x, Ak, Ak
t
)[M(Ak)h
1
(Ak
t
) M(Ak
t
)h
1
(Ak)] dk
t
=
_
R
3
s(x, Ak, w)[M(Ak)h
1
(w) M(w)h
1
(Ak)] dw
= (Q(h
1
))(Ak) = (Ak)
1
M(Ak) = k
1
M(k) = (Q(h
1
))(k),
78
and thus Q(h
1
A h
1
) = 0. Another computation yields (see (4.10))
_
R
3
h
1
(Ak)(k) dk =
_
R
3
_
R
3
s(x, k, k
t
)h
1
(Ak)M(k
t
) dk
t
dk
=
_
R
3
_
R
3
s(x, Ak, Ak
t
)h
1
(Ak)M(Ak
t
) dk
t
dk
=
_
R
3
_
R
3
s(x, v, w)h
1
(v)M(w) dwdv
=
_
R
3
h
1
(v)(v) dv
and
_
R
3
(h
1
A h
1
)dk = 0.
This is the orthogonality condition (4.12) which ensures the uniqueness of the solution of
Q(h
1
A h
1
) = 0. Therefore, h
1
A h
1
= 0. We conclude that h
1
remains invariant
under a rotation with axis k
1
. In particular, we can write
h
1
(k) = H
1
(k
1
, [k[
2
k
2
1
).
Now, let A be the isometric matrix of the linear mapping k (k
1
, k
2
, k
3
). Since
k k
1
M(k) is odd, a similar computation as above gives Q(h
1
A) = Q(h
1
) and
_
R
3
(h
1
A + h
1
)dk = 0.
This implies as above that h
1
A + h
1
= 0. Thus, h
1
is an odd function with respect to
k
1
.
In a similar way, we can show that
h
i
(k) = H
i
(k
i
, [k[
2
k
2
i
), i = 2, 3,
and H
i
are odd functions with respect to k
i
. In fact, all the functions H
i
equal H since,
for instance, exchanging k
1
and k
2
in
Q(H
1
(k
1
, k
2
2
+ k
2
3
)) = k
1
M(k
2
1
+ k
2
2
+ k
2
3
)
(with a slight abuse of notation) leads to
Q(H
1
(k
2
, k
2
1
+ k
2
3
)) = k
2
M(k
2
1
+k
2
2
+ k
2
3
) = Q(H
2
(k
2
, [k[
2
k
2
2
))
or Q(H
1
H
2
) = 0, and a similar argument as above shows that H
1
= H
2
. We set
H := H
1
.
79
Since H is odd with respect to the rst argument and [k[
2
k
2
j
does not depend on
k
j
, we obtain for all i ,= j,
_
R
3
k
i
h
j
(k) dk =
_
R
3
k
i
H(k
j
, [k[
2
k
2
j
) dk = 0. (4.17)
Furthermore,
_
R
3
k
i
h
i
(k) dk =
_
R
3
k
i
H(k
i
, [k[
2
k
2
i
) dk =
_
R
3
k
j
H(k
j
, [k[
2
k
2
j
) dk
=
_
R
3
k
j
h
j
(k) dk
for all i, j. This means that the integral is independent of i, and we can set
:=
_
R
3
k
1
h
1
(k) dk =
_
R
3
Q(h
1
)h
1
M
1
dk.
The parameter depends on x since h
1
depends on x through Q. Moreover, by (4.11),
(x) =
_
R
3
(Q(h
1
))h
1
M
1
dk 0
and by (4.17)
_
R
3
k h(k) dk = (x)I,
which shows (4.15).
Now, we are able to analyze the scaled Boltzmann equations (4.7)(4.8). Setting for-
mally = 0 gives
Q
n
(f
n
) = 0, Q
p
(f
p
) = 0.
By Lemma 4.1, these equations possess the solutions
f
n0
= n(x, t)M
n
, f
p0
= p(x, t)M
p
, (4.18)
respectively, where n(x, t) and p(x, t) are some parameters. Since
_
R
3
f
n0
dk = n(x, t),
_
R
3
f
p0
dk = p(x, t),
we can interpet n and p as scaled number densities of electrons and holes, repectively. In
order to obtain more information from (4.7)(4.8) we use the Hilbert method. The idea is
to expand f
n
and f
p
in terms of powers of ,
f
n
= f
n0
+ f
n1
+
2
f
n2
+ , (4.19)
f
p
= f
p0
+ f
p1
+
2
f
p2
+ , (4.20)
and to derive equations for f
ni
and f
pi
. We substitute this ansatz into (4.7)(4.8) and
equate coecients of equal powers of , noticing that Q
j
and I
j
are linear operators,
80
terms
0
:
Q
n
(f
n0
) = 0, Q
p
(f
p0
) = 0; (4.21)
terms
1
:
k
x
f
n0
E
e

k
f
n0
= Q
n
(f
n1
), (4.22)
k
x
f
p0
+ E
e

k
f
p0
= Q
p
(f
p1
); (4.23)
terms
2
:

t
f
n0
+ k
x
f
n1
E
e

k
f
n1
= Q
n
(f
n2
) +I
n
(f
n0
, f
p0
), (4.24)

t
f
p0
+ k
x
f
p1
+ E
e

k
f
p1
= Q
p
(f
p2
) + I
p
(f
n0
, f
p0
). (4.25)
Theorem 4.5 The Hilbert expansions (4.19)(4.20) are solvable up to second order if
and only if
n =
_
R
3
f
n0
dk, p =
_
R
3
f
p0
dk (4.26)
solve the (bipolar) drift-diusion equations

t
n div
x
J
n
= R(n, p), J
n
=
n
(
x
n +nE
e
), (4.27)

t
p div
x
J
p
= R(n, p), J
p
=
p
(
x
p pE
e
), (4.28)
R(n, p) = A(x)(np n
2
i
), (4.29)
where
A(x) =
1
n
2
i
_
R
3
_
R
3
g(x, k, k
t
) dk
t
dk, n
i
=
_
N
n
N
p
e
(EvEc)/2k
B
T
.
Proof: We already solved (4.21). With the solutions (4.18) we can rewrite (4.22)(4.23),
using
k
M
n
(k) = kM
n
(k),
k
M
p
(k) = (m

e
/m

h
)kM
p
(k),
Q
n
(f
n1
) = M
n
k (
x
n + nE
e
) = M
n
k
J
n

n
,
Q
p
(f
p1
) =
m

e
m

h
M
p
k (
x
p pE
e
) =
m

e
m

h
M
p
k
J
p

p
.
By Lemma 4.4, these equations have solutions and moreover, any solution can be expressed
as
f
n1
=
J
n

n
h
n
+
n
M
n
, f
p1
=
J
p

p
h
p
+
p
M
p
for some unspecied parameters
n
(x, t),
p
(x, t).
81
Equation (4.24) is solvable by Lemma 4.2 (1) if and only if
0 =
_
R
3
(
t
f
n0
+ k
x
f
n1
div
k
(E
e
f
n1
) I
n
(f
n0
, f
p0
)) dk
=
t
n +
_
R
3
k
x
f
n1
dk
_
R
3
I
n
(f
n0
, f
p0
) dk.
We infer from the denition (4.15) of
n
that
_
R
3
k
x
f
n1
dk =
_
R
3
k
x
_
J
n

n
h
n
+
n
M
n
_
dk
=
3

i,j=1

x
i
_
J
nj

n
_
R
3
k
i
h
nj
dk
_
+
x

_
R
3
kM
n
dk
=
3

i,j=1

x
i
J
nj

ij
= div
x
J
n
.
Furthermore,

_
R
3
I
n
(f
n0
, f
p0
) dk =
_
R
3
_
R
3
g(x, k
t
, k)
_
exp
_
E
c
E
v
k
B
T
+
[k[
2
2
+
m

e
m

h
[k
t
[
2
2
_
nM
n
(k)pM
p
(k
t
) 1
_
dk
t
dk
=
_
R
3
_
R
3
g(x, k, k
t
)
_
exp
_
E
c
E
v
k
B
T
_
np
N
n
N
p
1
_
dk
t
dk
=
_
R
3
_
R
3
g(x, k, k
t
)
_
np
n
2
i
1
_
dk
t
,
.
dk
= A(x)(np n
2
i
).
This shows (4.27) and (4.29). The proof of (4.28) is analogous.
The rst equations in (4.27) and (4.28) show that the quantities J
n
and J
p
can be
interpreted as particle current densities. They are the sum of the drift currents
n
nE
e
,

p
pE
e
and the diusion currents
n

x
n,
p

x
p, respectively, which explains the name
of the model.
Equations (4.27)(4.29) are in scaled form. In order to scale back to the physical
variables we notice that the scaled number densities, now called n
s
and p
s
, read
n
s
=
_
R
3
f
n0
dk
s
=
_

m

e
v
_
3
_
R
d
f
n0
dk =
_

2
m

e
k
B
T
_
3/2
n,
p
s
=
_
R
3
f
n0
dk
s
=
_

2
m

e
k
B
T
_
3/2
p,
82
where n and p are the unscaled variables. Thus, using
t
s
=
t

R
, E
e,s
=

0
U
T
E
e
, x
s
=
x

0
,
n,s
=
m

e
q
c

n
,
where U
T
= k
B
T/q is the thermal voltage, we obtain after some computations the unscaled
equations

t
n
1
q
div
x
J
n
= R(n, p), J
n
= q
n
(U
T

x
n + nE
e
), (4.30)

t
p +
1
q
div
x
J
p
= R(n, p), J
p
= q
p
(U
T

x
p pE
e
), (4.31)
the unscaled recombination-generation rate equals
R(n, p) = A(x)(np n
2
i
)
for some unscaled A(x) and with the intrinsic density
n
i
=
_
2k
B
T
_
m

e
m

2
_
3/2
exp
_
E
v
E
c
2k
B
T
_
which equals the intrinsic density (2.56) up to a factor 2/(2)
3/2
. The reason for this
dierence is that we ignored the moment density of states in (4.26) (see Section 2.4).
Equations (4.30)(4.31) are solved in x R
3
, t > 0, together with the initial conditions
n(x, 0) = n
I
(x), p(x, 0) = p
I
(x), x R
3
.
For a self-consistent treatment of the electric eld E
e
= V , equations (4.30)(4.31)
are supplemented by the Poisson equations

s
V = q(n p C(x)),
where
s
is the semiconductor permittivity and C(x) the doping concentration (see Propo-
sition 3.5).
Mathematically, (4.30)(4.31) are parabolic convection-diusion equations with diu-
sion coecients
D
n
:=
n
U
T
, D
p
:=
p
U
T
and so-called mobilities
n
and
p
. It is remarkable that the quotients of diusivities and
mobilities are constant; the equations
D
n

n
=
D
p

p
= U
T
are called Einstein relations.
The derivation of the drift-diusion model is mainly based on the following assump-
tions:
83
The mean free path
c
between two consecutive scattering events is much smaller
than the mean free path
R
between two recombination-generation events (typically,

c
10
8
m and
R
10
5
m).
The device diameter is of the order of
0
=

c
(typically,
0
10
6
m).
The electrostatic potential is of the order of U
T
= k
B
T/q (at room temperature:
U
T
0.026V).
This implies that the drift-diusion model is appropriate for semiconductor devices with
typical length scales not much smaller than 10
6
m and applied voltages much smaller
than 1V. However, in applications, this model is used also for voltages much larger than
U
T
, sometimes together with some correction forms.
4.2 Derivation of the hydrodynamic equations
In the previous section we have derived a uiddynamical model from the Boltzmann
equation by using the Hilbert expansion method. A second approach for the derivation
of uid models is the moment method which consists in deriving evolution equations for
averaged quantities, like particle density, current density, and energy.
We consider the semi-classical Boltzmann equation in the parabolic-band approxima-
tion

t
f +

m

e
k
x
f
q

E
e

k
f = Q(f), x, k R
d
, t > 0. (4.32)
We assume that the collision operator splits into two parts
Q(f) = Q
1
(f) +Q
2
(f),
with the mean free paths
c
for the rst part and for the second part. Similarly as in
Section 4.1 we choose the reference length , the reference velocity v =
_
k
B
T
L
/m

e
, the
reference time = /v, the reference wave vector m

e
v/, and the reference eld U
T
/,
where U
T
= k
B
T
L
/q, and T
L
is the lattice temperature. Then, dening
x = x
s
, t = t
s
, k =
m

e
v

k
s
, E
e
=
U
T

E
e,s
and
Q
1
(f) =
1

c
Q
1,s
(f), Q
2
(f) =
1

Q
2,n
(f),
where
c
=
c
/v, inserting this scaling into (4.32), multiplying this equation by , and
omitting the index s, gives the scaled equation

t
f

+ k
x
f

E
e

k
f

=
1

Q
1
(f

) + Q
2
(f

), (4.33)
84
where =
c
/.
For the following we set
g(k)) =
_
R
d
g(k) dk
for any function g depending on k. We call the averaged quantities (k)f) moments of
f, In particular, f), k
i
f), and
1
2
[k[
2
f) are called the zeroth, rst, and second moments
of f, respectively.
The collision operators are assumed to satisfy

1
Q
1
(f) + Q
2
(f)) = 0, (4.34)
k(
1
Q
1
(f) + Q
2
(f))) = kf), (4.35)

1
2
[k[
2
(
1
Q
1
(f) + Q
2
(f))) =
d
2
f) c
1
2
[k[
2
f) (4.36)
for all functions f. The following example shows that the relaxation-time operator (3.40)
fullls these conditions.
Example 4.6 Let
(Q
1
(f))(x, k, t) = f(x, , t))M(k), (Q
2
(f))(x, k, t) = f(x, k, t),
where M(k) = (2)
d/2
e
[k[
2
/2
. Then
1
Q
1
+ Q
2
satises (4.34)(4.36). Indeed, since
M) =
1
(2)
d/2
_
R
d
e
[k[
2
/2
dk = 1,
k
i
M) =
1
(2)
d/2
_
R
d
k
i
e
k
2
i
/2
dk
i

j,=i
_
R
e
k
2
j
/2
dk
j
= 0,

1
2
[k[
2
M) =
1
2(2)
d/2
d

i=1
_
R
d
k
2
i
e
k
2
i
/2
dk
i

j,=i
_
R
e
k
2
j
/2
dk
j
,
=
1
2(2)
d/2
d

i=1
(2)
1/2
(2)
(d1)/2
=
d
2
,
we obtain

1
Q
1
(f) +Q
2
(f)) = f)M) f) = 0,
k(
1
Q
1
(f) + Q
2
(f))) = f)kM) kf) = kf),

1
2
[k[
2
(
1
Q
1
(f) + Q
2
(f))) = f)
1
2
[k[
2
M)
1
2
[k[
2
f) =
d
2
f)
1
2
[k[
2
f).

Multiplying the scaled Boltzmann equation (4.33) by 1, k,
1
2
[k[
2
, respectively, and
integrating over k R
d
yields the so-called moment equations

t
f

) + div
x
kf

) = 0 (4.37)

t
kf

) + div
x
k kf

) +E
e
f

) = kf

) (4.38)

1
2
[k[
2
f

) + div
x

1
2
k[k[
2
f

) + E
e
kf

) =
d
2
kf

)
1
2
[k[
2
f

), (4.39)
85
where we have used the divergence theorem to compute the force terms:
E
e

k
f

) = div
k
(E
e
f

)) = 0,
k
i
E
e

k
f

) =
k
k
i
E
e
f

) = E
e,i
f

),

1
2
[k[
2
E
e

k
f

) =
1
2

k
[k[
2
E
e
f

) = k E
e
f

).
The goal of the moment method is to express all integrals in the moment equations in
terms of the rst moments f

),kf

) and
1
2
[k[
2
f

). However, it is not clear how to express


the uxes kkf

) and
1
2
k[k[
2
f

) in terms of the rst moments. Moreover, the third-order


moment ux
1
2
k[k[
2
f

) cannot be written with the help of the moments up to second


order. Therefore, an additional assumption is necessary. We present two approaches. The
rst one consists in assuming that the mean free path
c
of the rst part of the collision
operator is much smaller than the typical device length such that =
c
/ 1 and to
perform the formal limit 0. This idea is due to Bardos, Golse and Levermore [6, 7]
(also see the review [36]). The second approach closes the moment hierarchy by choosing
a special! function f

which gives the minimum of the entropy functional. This approach


is called the entropy minimization principle and has been used by Levermore [42, 43].
For the rst approach we impose the following assumptions. Let f

be a solution to
(4.33) such that, as 0,
f

f pointwise almost everywhere, (4.40)


(k)f

) (k)f) for (k) = 1, k


1
, . . . , k
d
, [k[
2
, (4.41)
and assume that
Q
1
(f) = 0 if and only if f = exp( + k + [k[
2
), (4.42)
where , R and = (
1
, . . . ,
d
)

R
d
.
The last condition implies that there exist functions n(x, t) > 0, u(x, t) R, and
T(x, t) > 0 such that f N(Q
1
) can be written equivalently as a so-called shifted
Maxwellian (see (3.41))
M(x, k, t) =
n(x, t)
(2T(x, t))
d/2
exp
_

[k u(x, t)[
2
2T(x, t)
_
. (4.43)
The functions n, u, and T can be interpreted physically as the particle density, mean
velocity, and particle temperature, respectively.
Lemma 4.7 The Maxwellian (4.43) satises
M) = n, kM) = nu, k kM) = n(u u) + nTI,
86

1
2
[k[
2
M) = n(
1
2
[u[
2
+
d
2
T), k[k[
2
M) = 2nu(
1
2
[u[
2
+
d+2
2
T),
where I is the identity matrix in R
dd
and a b = a

b for a, b R
d
.
The quantity J = nu is called the particle current density, and e =
1
2
[u[
2
+
d
2
T is the
total energy consisting of the sum of kinetic and thermal energy.
Proof: We use the identities
_
R
x
2m
e
x
2
/2
dx = (2m1)(2m3) . . . 5 3 1

2,
_
R
x
2m+1
e
x
2
/2
dx = 0 (4.44)
for m N
0
and the transformation z = (k u)/

T to compute
M) = n
_
R
d
e
[z[
2
/2
dz = n,
kM) =
n
(2)
d/2
_
R
d
(u +

Tz)e
[z[
2
/2
dz = nu
k kM) =
n
(2)
d/2
_
R
d
(u +

Tz) (u +

Tz)e
[z[
2
/2
dz
= n(u u) +nT(2)
d/2
_
R
d
z ze
[z[
2
/2
dz
= n(u u) +nTI.
From the last identity follows

1
2
[k[
2
M) =
1
2

d
i=1
(nu
2
i
+ nT) = n(
1
2
[u[
2
+
d
2
T),
and nally,
k
i
[k[
2
M) =
n
(2)
d/2
d

j=1
_
R
d
(u
i
+

Tz
i
)(u
j
+

Tz
j
)
2
e
[z[
2
/2
dz
= nu
i
[u[
2
+
u
i
nT
(2)
d/2
d

j=1
z
2
j
e
[z[
2
/2
dz
+2
nT
(2)
d/2
d

j=1
u
j
_
R
d
z
i
z
j
e
[z[
2
/2
dz
+
nT
3/2
(2)
d/2
_
R
d
z
i
[z[
2
e
[z[
2
/2
dz
= nu
i
[u[
2
+ nT
d

j=1
(u
i
+ 2u
j

ij
)
= 2nu
i
_
1
2
[u[
2
+
d + 2
2
T
_
.

87
Theorem 4.8 Let the assumptions (4.40)(4.42) hold. Then, as 0, the moment
equations (4.37)(4.38) become

t
n divJ = 0, (4.45)

t
J div
_
J J
n
_
(nT) + nE
e
= J, (4.46)

t
(ne) div[J(e + T)] + J E
e
= n(e
d
2
), (4.47)
where J = nu and e =
1
2
[u[
2
+
d
2
T.
Equations (4.45)(4.47) are called the (scaled) hydrodynamic equations for semicon-
ductors. They have the same form as the Euler equations of gas dynamics subject to the
electric force E
e
and the momentum and energy relaxation terms J, n(e
d
2
), re-
spectively. Mathematically, they are nonlinear equations of hyperbolic type. For constant
particle temperature T = 1 (or T = T
L
in unscaled variable) equations (4.45)(4.46) are
referred to as the isothermal hydrodynamic model. For Coulomb forces E
e
= V , they
are supplemented with the scaled Poisson equation

2
D
V = n C(x),
where the so-called Debye length
2
D
=
s
U
T
/(qL
2
max C) comes from the scaling. Equa-
tions (4.45)(4.46) are solved for x R
d
, t > 0, together with the initial conditions
n(x, 0) = n
I
(x), J(x, 0) = J
I
(x), T(x, 0) = T
I
(x), x R
d
,
which determine the initial function e(x, 0).
Proof: Assumption (4.40) implies that for all smooth functions ,
_
R
2d+1
(
t
f

+k
x
f

E
e

k
f

)dk dx =
_
R
2d+1
f

(
t
+k
x
E
e

k
) dk dx
is bounded; therefore, in the weak sense,
Q
1
(f) = lim
0
Q
1
(f

) = 0
and by assumption (4.42), f = M. Assumption (4.41) allows us to pass to the limit 0
in the moment equations (4.37)(4.38) yielding

t
M) + divkM) = 0,

t
kM) + divk kM) + E
e
M) = kM),

1
2
[k[
2
M) + div
1
2
k[k[
2
M) + E
e
kM) =
d
2
M)
1
2
[k[
2
M).
Then Lemma 4.7 gives the assertion.
88
Remark 4.9 The unscaled hydrodynamic equations can be written as follows:

t
n
1
q
divJ = 0 (4.48)

t
J
1
q
div
_
J J
n
_
+
qk
B
m

e
(nT)
q
2
m

e
nE
e
=
J

(4.49)

t
(ne)
1
q
div[J(e + k
B
T)] J E
e
=
n

_
e
d
2
k
B
T
L
_
, (4.50)
where the energy is given by
e =
m

e
2q
2
[J[
2
n
2
+
d
2
k
B
T.
We now turn to the second approach of deriving the hydrodynamic model. The main
hypotheses of the rst approach are 1 and that the kernel of Q
1
consists of the
shifted Maxwellians such that they can be used as a closure function in the moment
equations. The idea of the second approach is to choose a closure function f

in the
moment equations, which minimizes the entropy functional
(H(f))(x, t) =
_
R
d
f(log f 1) dk. (4.51)
We will see that f

is in fact equal to the shifted Maxwellian. We present this entropy


minimization method in a slightly more general framework.
Let
i
(k) be some monomials in the variables k
1
, . . . , k
d
for i = 0, . . . , N and set = 1.
The moment equations are now

i
(k)f) + div
x
k
i
(k)f) E
e

i
(k)
k
f) =
i
(k)(Q
1
+Q
2
)) (4.52)
for x R
d
, t > 0, i = 0, . . . , N. In order to close this system of equations we choose the
distribution function f

which minimizes the entropy (4.51) among those functions whose


moments m(x) = (m
0
(x), . . . , m
N
(x)) are given. In the rst approach, we have used the
moments m
0
= n, m
i
= nu
i
(i = 1, . . . , d), and m
d+1
= e. More precisely, let f

be the
solution to the so-called Gibbs problem
H(f

) = minH(f) :
i
(k)f) = m
i
(x) for all x, i. (4.53)
Notice that the physical entropy is H(f), i.e., the physical entropy has to be maximized.
Lemma 4.10 The formal solution to (4.53) is
f

(x, k) = e
(x)(k)
, (4.54)
where (x) are Lagrange multipliers and (k) = (
0
(k), . . . ,
N
(k))

.
89
Proof: Using Lagrange multipliers, we have to analyze the functional
G(f, ) =
_
R
d
f(log f 1) dk +
N

i=0

i
(
i
f) m
i
).
We claim that
G(f, )
f
(g) =
_
R
d
g log f dk + g). (4.55)
Indeed, we compute formally
1

[G(f +g, ) G(f, )] =


_
R
d
_
1

f log
_
1 +
g
f
_
+ g(log(f + g) 1)
_
dk +
1

g).
Since log(1 +g/f) g/f ( 0), we obtain
lim
0
1

[G(f + g, ) G(f, )] = lim


0
_
R
d
g log(f + g) dk + g)
=
_
R
d
g log f dk + g),
which proves (4.55).
Now, the necessary condition
G(f

, )
f
(g) = 0 for all functions g
for an extremal f leads to
_
R
d
g(log f

+ ) dk = 0 for all functions g


and hence, log f

= .
The Lagrange multipliers
i
(x) can be derived by inserting (4.54) into the constraints
(4.53):
m(x) =
_
R
d

i
(k)e
(x)(k)
dk for x R
d
, i = 0, . . . , N. (4.56)
These are implicit equations for
0
, . . . ,
N
which, generally, cannot be directly inverted
to give explicit expressions for
0
, . . . ,
N
. However, when choosing
m = (n, nu, ne) and (k) =
_
1, k,
1
2
[k[
2
_
, (4.57)
we can solve (4.56). To see this, we observe that
f

= exp( ) = exp
_

j=1

j
k
j

1
2

d+1
[k[
2
_
90
can be equivalently written as the shifted Maxwellians
f

=
n
(2T)
d/2
exp
_

[k u[
2
2T
_
,
where now the parameters n, u, T play the role of the Lagrange multipliers. The notation
is chosen correctly since the moments of this Maxwellian yield the moments (n, nu, ne).
This shows that (4.56) is satised. Then Lemma 4.7 and (4.34)(4.36) show that the choice
f = f

in the moment equations (4.52) yields the hydrodynamic model (4.45)(4.47). We


have shown the following result.
Theorem 4.11 Assume that (4.34)(4.36) hold and let f

be the solution to the Gibbs


problem (4.53). Then, choosing m
i
(x) and
i
(k) according to (4.57), the moments (n, nu, e)
solve the hydrodynamic equations (4.45)(4.47).
The hydrodynamic equations have been rst introduced by Bltekjr [10] and Bac-
carani and Wordeman [5] in the context of semiconductor devices. In their model, some-
times called the BBW equations, the heat conduction term div((n, T)T) is added to
the left-hand side of (4.50). The heat conductivity (n, T) is usually modeled in R
3
by
(n, T) =
3
2
k
2
B

e
nT
(see [5]). The heat conduction term makes equation (4.50) parabolic since we can write it
as
d
2
k
B

t
T div((n, T)T) = function of n, J, T and rst derivatives.
In particular, the heat conduction term is often used in numerical simulations. However,
it is not clear how to justify this term.
4.3 Derivation of the spherical harmonics expansion equations
In Section 4.1 we have derived the drift-diusion model from the Boltzmann equation by
the Hilbert expansion method under the assumption that the mean free path between
two consecutive collision events is much smaller than the typical time. In this section we
perform a Hilbert expansion using the same hypothesis but a dierent collision operator.
We show that this approximation leads to the Spherical Harmonics Expansion (SHE)
model.
We consider the semi-classical Boltzmann equation

t
f + v(k)
k
f
q

E
e

k
f = Q
el
(f) + Q
inel
(f), x R
3
, k B, t > 0, (4.58)
91
where v(k) =
1

k
E(k), Q
el
(f) models elastic collisions and is given by
(Q
el
(f))(x, k, t) =
_
B
(x, k, k
t
)(E(k
t
) E(k))(f
t
f) dk
t
, (4.59)
with the scattering rate (x, k, k
t
) and f = f(x, k, t), f
t
= f(x, k
t
, t) (see Example 3.6).
Furthermore, Q
inel
(f) models inelastic collisions (e.g., inelastic corrections to electron-
phonon and electron-electron scattering) and will not specied in this section. We impose
the following assumptions:
The operator Q
inel
is linear.
The scattering rate is positive and symmetric in k and k
t
:
(x, k, k
t
) > 0, (x, k, k
t
) = (x, k
t
, k) for all x R
3
, k, k
t
B. (4.60)
We scale the Boltzmann equation (4.58) by using the same scaling as in Section 4.1,
i.e., we x the reference velocity v
0
=
_
k
B
T/m

e
, the characteristic times
el
,
inel
of elastic
and inelastic collisions, respectively, with corresonding mean free paths
el
=
el
v
0
and

inel
=
inel
v
0
, the reference length
0
=

el

inel
and time
0
=
0
/v
0
. Then, with the
scaling
x =
0
x
s
, t =
inel
t
s
, k =
m

e
v
0

k
s
, E
e
=
U
T

0
E
e,s
,
E(k) = m

e
v
2
0
E
s
(k
s
), Q
el
(f) =
1

el
Q
el,s
(f), Q
inel
(f) =
1

inel
Q
inel
(f),
we obtain, omitting the index s and introducing
2
=
el
/
inel
,

t
f + (
k
E(k)
x
f E
e

k
f) = Q
el
(f) +
2
Q
inel
(f), (4.61)
since v
0

el
/
0
= . We assume that 1.
We now analyze the collision operator Q
el
. In [8] the following results are shown.
Lemma 4.12 Assume that (4.60) holds. Then
(1) Q
el
is self-adjoint and non-negative on L
2
(B), i.e.

_
B
Q
el
(f)f dk 0 for all f L
2
(B).
(2) The kernel of Q
el
is given by
N(Q
el
) = f L
2
(B) : f only depends on E(k).
92
(3) The equation Q
el
(f) = h has a solution if and only if
_
B
h(k)(e E(k)) dk = 0 for all e R(E). (4.62)
(4) For all functions g only depending on E(k) and all f L
2
(B) there holds
Q
el
(gf) = gQ
el
(f). (4.63)
In the statement of the lemma we have omitted some technical assumptions on the
regularity of (x, k, k
t
) and g(E(k)); we refer to [8] for details.
Proof: (1) Let x R
3
, f, g L
2
(B), and set f = f(k), f
t
= f(k
t
), g = g(k), g
t
= g(k
t
),
E = E(k), and E
t
= E(k
t
). Then, using the property
_
B
(E
t
E)(k, k
t
) dk
t
=
_
B
(E E
t
)(k, k
t
) dk
t
for any (smooth) function and employing the symmetry of , we compute
_
B
Q
el
(f)g dk =
1
2
_
B
2
(x, k, k
t
)(E
t
E)(f
t
f)g dk
t
dk
+
1
2
_
B
2
(x, k
t
, k)(E E
t
)(f f
t
)g
t
dk dk
t
=
1
2
_
B
2
(x, k, k
t
)(E
t
E)(f
t
f)(g
t
g) dk
t
dk
=
_
B
Q
el
(g)f dk.
This shows that Q
el
is self-adjoint. Taking f = g gives

_
B
Q
el
(f)f dk =
1
2
_
B
2
(x, k, k
t
)(E
t
E)(f
t
f)
2
dk
t
dk 0 (4.64)
and thus, Q
el
is non-negative.
(2) Let f N(Q
el
). Then (4.64) implies that
(E(k
t
) E(k))(f(k
t
) f(k))
2
= 0 for k, k
t
B.
Heuristically, (x) = 0 for x ,= 0 and (x) ,= 0 for x = 0. This motivates
E(k
t
) = E(k) and f(k
t
) = f(k) for k, k
t
B.
Hence, f must be constant on each energy surface k B : E(k) = e. We infer that f
is a function of E(k) only.
93
(3) We show that N(Q
el
)

only consists of functions satisfying (4.62). Then the self-


adjointness of Q
el
and Fredholms alternative give the conclusion. Let h N(Q
el
)

and
f N(Q
el
). By part (2), we can write f(k) = g(E(k)) for some function g and
0 =
_
B
h(k)f(k) dk =
_
B
h(k)g(E(k)) dk =
_
B
h(k)
_
R(E)
g(e)(E(k) e) de dk
=
_
R(E)
__
B
h(k)(E(k) e) dk
_
g(e) de.
This equation holds for all f N(Q
el
) and thus for any function g. Hence
_
B
h(k)(E(k) e) dk = 0 for e R(E).
Thus, functions of N(Q
el
)

are satisfying (4.62). On the other hand, if h fullls (4.62) the


same arguments as above show that h N(Q
el
)

.
(4) We claim that
_
B
(x, k, k
t
)(E
t
E)f
t
g(E
t
) dk
t
= g(E(k))
_
B
(x, k, k
t
)(E
t
E)f
t
dk
t
. (4.65)
Indeed, by the coarea formula (Theorem 2.15) we can write formally
_
B
(x, k, k
t
)(E
t
E)f
t
g(E
t
) dk
t
=
_
R(E)
_
E
1
(e)
(x, k, k
t
)(e E(k))f(k
t
)g(e)
dF(k
t
)
[
k
E(k
t
)[
de
=
_
R(E)
_
E
1
(e)
(x, k, k
t
)(e E(k))f(k
t
)g(E(k))
dF(k
t
)
[
k
E(k
t
)[
de
= g(E(k))
_
B
(x, k, k
t
)(E
t
E)f
t
dk
t
.
This gives
(Q
el
(gf))(k) =
_
B
(x, k, k
t
)(E
t
E)f
t
g(E
t
) dk
t

_
B
(x, k, k
t
)(E
t
E)fg(E) dk
t
= g(E(k))
_
B
(x, k, k
t
)(E
t
E)f
t
dk
t
g(E(k))
_
B
(x, k, k
t
)(E
t
E)f dk
t
= g(E(k))(Q
el
(f))(k).
This nishes the proof of the lemma.
94
Remark 4.13 From the proof of Lemma 4.12 (4) we conclude the following general result:
For all functions (k, k
t
) and g(E(k)) it holds
_
B
(k, k
t
)(E
t
E)g(E
t
) dk
t
= g(E(k))
_
B
(k, k
t
)(E
t
E) dk
t
. (4.66)
Now we insert the Hilbert expansion.
f = f
0
+ f
1
+
2
f + (4.67)
into the scaled Boltzmann equation (4.61), use the linearity of the collision operators Q
el
and Q
inel
and identify terms of equal order in to nd
terms
0
:
Q
el
(f
0
) = 0; (4.68)
terms
1
:
Q
el
(f
1
) =
k
E(k)
x
f
0
E
e

k
f
0
; (4.69)
terms
2
:
Q
el
(f
2
) =
t
f
0
+
k
E(k)
x
f
1
E
e

k
f
1
Q
inel
(f
0
). (4.70)
These equations can be solved using the properties of the operator Q
el
.
Theorem 4.14 The Hilbert expansion (4.67) is solvable up to second order if and only
if f
0
(x, k, t) = F(x, E(k), t) and F is a solution of
N()
t
F div
x
J + E
e

J

= S(F), x R
3
, R(E), t > 0, (4.71)
where
J(x, , t) = D(x, )
_

x
F E
e
F

_
(4.72)
is the current density,
N() =
_
B
( E(k)) dk (4.73)
is called the density of states of energy (compare with Lemma 2.14),
D(x, ) =
_
B

k
E(k) (x, k)( E(k)) dk R
33
(4.74)
is the diusion matrix, where is a solution of Q
el
() =
k
E(k), and
(S(F))(x, , t) =
_
B
(Q
inel
(F))(x, k, t)( E(k)) dk (4.75)
is the collision operator, averaged over the energy surface = E(k).
95
Equations (4.71)(4.72) are called the Spherical Harmonic Expansion (SHE) model.
It has been rst derived in the physical literature for spherically symmetric band dia-
grams and rotationally invariant collision operators as a numerical approximation of the
Boltzmann equation [34, 35] or through a macroscopic limit [28] which explains the name.
Notice, however, that the above SHE model is derived without the assumption of spherical
symmetric energy bands. The presented expansion is no more spherical but rather on
surfaces of constant energy. The SHE model is supplemented by the initial conditions
F(x, , 0) = F
I
(x, ), x R
3
, R(E),
and appropriate boundary conditions for which we do not detail here.
Proof: By Lemma 4.12 (2), any solution to (4.68) can be written as
f
0
(x, k, t) = F(x, E(k), t)
for some function F. Thus we can reformulate (4.69) as
Q
el
(f
1
) =
k
E
_

x
F E
e
F

_
.
We claim that any solution of this equation can be written as
f
1
(x, k, t) = (x, k)
_

x
F E
e
F

_
+ F
1
(x, E(k), t), (4.76)
where F
1
N(Q
el
) and (x, k) is a solution of
Q
el
() =
k
E. (4.77)
Notice that depends on the parameter x since in the denition of Q
el
also depends
on x. By Lemma 4.12 (3), (4.77) is solvable if and only if
_
B

k
E(e E(k)) dk = 0 for all e R(E).
To see this we introduce the Heaviside function H, dened by H(x) = 0 for x < 0 and
H(x) = 1 for x > 0. This function has the property that H
t
(x) = (x). Hence
_
B

k
E(e E(k)) dk =
_
B

k
H(e E(k)) dk = 0, (4.78)
since E(k) is periodic on B. Thus (4.77) is solvable. Observing that
x
F E
e
(F/)
only depends on E(k) (and on the parameters x, t), Lemma 4.12 (4) shows that
Q
el
(f
1
) =
_

x
F E
e
F

_

k
E = Q
el
_

x
F E
e
F

__
.
96
The linearity of Q
el
implies that
Q
el
_
f
1
+
_

x
F E
e
F

__
= 0
and therefore, (4.76) follows. In the following, we take a particular solution to (4.69) by
choosing F
1
= 0.
It remains to solve (4.70). Again, by Lemma 4.12 (3), this operator equation is solvable
if and only if
_
B
(
t
f
0
+
k
E
x
f
1
E
e

k
f
1
Q
inel
(f
0
)) (e E(k)) dk = 0 (4.79)
for all e R(E). We claim that this condition is equivalent to (4.71). To show this, we
compute the terms separately. By (4.66) we obtain for the rst term
_
B

t
f
0
(e E(k)) dk =
_
B

t
F(x, E(k), t)(e E(k)) dk
=
t
F(x, e, t)
_
B
(e E(k)) dk = N(e)
t
F(x, e, t).
We introduce the electron current density by
J(x, e, t) =
_
B

k
E(k)f
1
(x, k, t)(e E(k)) dk.
Then the second term in (4.79) becomes
_
B

k
E
x
f
1
(e E(k)) dk = div
x
J(x, e, t).
The current density can be reformulated in terms of the function F by using the expression
(4.76) for f
1
and the property (4.66):
J(x, e, t) =
_
B

k
E(k)(x, k)
_

x
F E
e
F

_
(x, E(k), t)(e E(k)) dk
= D(x, e)
_

x
F E
e
F

_
(x, e, t).
For the computation of the third term in (4.79) we choose a smooth function with
compact support on R(E) and use the denition of the distribution and integration by
parts to obtain
_
R(E)
(e)
_
B

k
f
1
(x, k, t)(e E(k)) dk de
=
_
B
(E(k))
k
f
1
(x, k, t) dk =
_
B
f
1
(x, k, t)
t
(E(k))
k
E(k) dk
=
_
R(E)

t
(e)
_
B

k
E(k)f
1
(x, k, t)(e E(k)) dk de
=
_
R(E)

t
(e)J(x, e, t) de =
_
R(E)
(e)
J

(x, e, t) de.
97
Since is arbitrary, it follows
_
B

k
f
1
(x, k, t)(e E(k)) dk =
J

(x, e, t).
We have shown that (4.79) is equivalent to (4.71)(4.72).
The advantage of the SHE model is twofold:
The SHE equations have to be solved in 3 + 1 dimensions of the position-energy
space instead of the (3 + 3)-dimensional phase space of the Boltzmann equation.
The SHE model is of parabolic type which simplies its numerical solution.
In order to see that the SHE model is of parabolic type, we assume that the electric
eld E
e
is a gradient, i.e, E
e
=
x
V , and we introduce the total energy variable
u = V (x, t).
Then, with the change of unknowns
F(x, , t) = f(x, V (x, t), t) = f(x, u, t),
D(x, ) = d(x, V (x, t), t) = d(x, u, t),
N() = N(u + V (x, t)) = (x, u, t),
we obtain

x
F =
x
f
x
V
f
u
=
x
f
x
V
F

or

F :=
_

x
+
x
V

_
F =
x
f
and

t
F =
t
f
t
V
f
u
.
Since we can write (4.71)(4.72) equivalently as
N()
t
F

J = S(F), J = D

F,
we obtain, in the total energy variable,
(x, u, t)
t
f div
x
(d(x, u, t)
x
f) = S(F) + (x, u, t)
t
V
f
u
.
The parabolicity now follows from the positive (semi-) deniteness of the diusion matrix
d or D, respectively, as stated in the following lemma.
Lemma 4.15 The diusion matrix D(x, ) R
33
is symmetric and positive semi-
denite.
98
In [8, Prop. 3.6] a stronger property for D(x, ) is shown: There exists K > 0 such
that
D
ij
(x, )
K
N()
_
B
E
k
i
(k)
E
k
j
(k)( E(k)) dk, i, j = 1, 2, 3. (4.80)
In particular, det D = 0 at critical points of E, (i.e.
k
E(k) = 0 and = E(k)).
Proof: We use the symmetry of Q
el
(Lemma 4.12 (1)) and the property (4.63) of Lemma 4.12
(4) to show the symmetry of D(x, ). Let be a (smooth) function. Then
_
R
D
ij
(x, )() d =
_
R
_
B
E
k
i

j
( E(k)) dk () d
=
_
B
Q
el
(
i
)
j
(E(k)) dk =
_
B
Q
el
((E(k))
i
)
j
dk
=
_
B
Q
el
((E(k))
j
)
i
dk =
_
R
D
ji
(x, )() d.
since is arbitrary, D
ij
(x, ) = D
ji
(x, ) for all x R
3
, R.
In order to show that D(x, ) is positive semi-denite, let z R
3
. Then, by denition
of Q
el
,
z

D(x, )z =
3

i,j=1
_
B
z
i
Q
el
(
i
)
j
z
j
( E(k)) dk
=
3

i,j=1
_
B
_
B
(x, k, k
t
)(E(k
t
) E(k))z
i
(
i
(k
t
)
i
(k))z
j

j
(k)
( E(k)) dk
t
dk.
Similar as in the proof of Lemma 4.12 (1), we can write
z

D(x, )z =
1
2
3

i,j=1
_
B
_
B
(x, k, k
t
)(E(k
t
) E(k))( E(k))
z
i
(
i
(k
t
)
i
(k))z
j
(
j
(k
t
)
j
(k)) dk
t
dk
=
1
2
_
B
_
B
(x, k, k
t
)(E(k
t
) E(k))( E(k))

_
3

i=1
z
i
(
i
(k
t
)
j
(k))
_
2
dk
t
dk
0.
For the last equality we have used the elementary identity
n

i,j=1
a
i
a
j
=
_
n

i=1
a
i
_
2
for all a
n
, . . . , a
n
R.
99
The lemma is shown.
More explicit expressions for the density of states N() and the diusion matrix D(x, )
can be derived for special energy bands (see [8, Sec. 3.4]).
Example 4.16 (Spherical symmetric energy bands)
Assume that the scattering rate only depends on E(k), i.e.
(x, k, k
t
) = s(x, E(k)) for all k, k
t
with E(k) = E(k
t
),
and that E(k) is spherical symmetric, i.e. E = E([k[), and strictly monotone in [k[, i.e.,
there exists a function : R R such that
[k[
2
= (E([k[)).
Notice that the rst assumption makes sense since due to the term (E(k
t
) E(k)) in
the denition of Q
el
, the scattering rate needs to be dened only on the surface k
t

B : E(k
t
) = E(k) of energy E(k). The second assumption implies that we can choose
B = R
3
.
We claim that
N() = 2
_
()
t
(), (4.81)
D(x, ) =
4
3
()
s(x, )
t
()
2
I, (4.82)
where I R
33
is the identity matrix.
First we reformulate D(x, ). From the rst assumption and the denition (4.73) of
the density of states follows, with E = E(k), E
t
= E(k
t
),
(Q
el
(f))(x, k, t) = s(x, E(k))
_
R
3
(E
t
E)f(k
t
) dk
t
s(x, E(k))N(E(k))f(k)
=
([f] f)(x, k, t)
(x, E(k))
, (4.83)
where (x, ) = 1/s(x, )N() is called relaxation time and
[f](k) =
1
N(E(k))
_
R
3
(E
t
E)f(k
t
) dk
t
is the average of f on the energy surface k
t
: E(k
t
) = E(k). The expression (4.83) is
called relaxation-time operator (compare with (3.40)).
We claim that the solution of Q
el
() =
k
E can now be written explicitly as
(x, k) = (x, E(k))
k
E(k).
100
Indeed, by (4.64) and (4.78),
[](k) =
1
N(E(k))
_
R
3
(E
t
E)(x, E
t
)
k
E(k
t
) dk
t
=
(x, E(k))
N(E(k))
_
R
3
(E
t
E)
k
E(k
t
) dk
t
= 0,
and hence
Q
el
() =
[]
(x, E(k))
=
k
E(k).
Thus, we can write
D(x, ) =
_
R
3

k
E(k)
k
E(k)(x, E(k))( E(k)) dk
= (x, )
_
R
3

k
E(k)
k
E(k)( E(k)) dk.
This shows that the estimate (4.80) is sharp.
The above expression for D(x, ) can be further simplied under the second assump-
tion. As E only depends on [k[, it is convenient to use spherical coordinates k = , where
> 0 and
=
_
_
_
sin cos
sin sin
cos
_
_
_
, 0 < 2, 0 < .
Then, transforming = E() with d = E
t
() d,
D(x, ) = (x, )
_

0
_
2
0
_

0
E
t
() E
t
()( E())
2
sin d dd
= (x, )
_
R
E
t
()( )
2
d
_
2
0
_

0
sin d d.
By the denition of the distribution, the rst integral is
_
R
E
t
()( )
2
d = E
t
([k[)[k[
2
with = E([k[).
For the second integral an elementary computation gives
_
2
0
_

0
sin d d =
4
3
I,
where I is the identity matrix. Therefore
D(x, ) =
4
3
(x, )E
t
([k[)[k[
2
with = E([k[).
101
Dierentiation of [k[
2
= (E([k[)) with respect to [k[ yields 2[k[ =
t
(E([k[))E
t
([k[) and
therefore
E
t
([k[) =
2
_
()

t
()
with = E([k[). (4.84)
We conclude that
D(x, ) =
8
3
(x, )()
3/2

t
()
I =
8
3
()
3/2
s(x, )N()
t
()
I.
The density of states can be computed as
N() =
_
R
3
( E([k[)) dk =
_

0
_
2
0
_

0
( E())
2
sin d dd
= 4
_
R
( )E
t
()
1

2
d =
4[k[
2
E
t
([k[)
with = E([k[).
Employing (4.84) we infer that
N() = 2
_
()
t
()
and
D(x, ) =
4
3
()
s(x, )
t
()
2
I.
This shows (4.81) and (4.82).
Example 4.17 (Parabolic band approximation)
We assume that (x, k, k
t
) = s(x, E(k)) for all k, k
t
R
3
such that E(k) = E(k
t
) and
E(k) =
1
2
[k[
2
.
Then () = 2 and
N() = 4

2, D(x, ) =
2
3s(x, )
I.
With the specication (see [24, (2.6)] or [45, (3.63)])
(x, ) =
1
s
0

N()
, > 1, (4.85)
we obtain s(x, ) = 1/(x, )N() = s
0

and hence
D(x, ) =
2
3s
0

1
I.
102
It remains to determine the average collision operator S(F) (see (4.75)). The precise
structure of this term depends on the assumptions on the inelastic collision operator
Q
inel
. Here we only notice that a simplied expression is given by the Fokker-Planck
approximation [60]
S(F) =

_
s
1

N()
2
_
F +
F

__
. (4.86)
Remark 4.18 The SHE equations read in unscaled form
N()
t
F
1
q
div
x
J E
e

J

= S(F),
J(x, , t) = D(x, )
_

x
F qE
e
F

_
,
where
N() =
_
B
( E(k)) dk,
D(x, ) =
q

_
B

k
E(k) (x, k)( E(k)) dk.

4.4 Derivation of the energy-transport equations


The SHE model of Section 4.3 is much simpler than the semi-classical Boltzmann equation
since it has a parabolic structure and has to be solved in the one-dimensional energy
variable instead of the three-dimensional wave vector. However, it is numerically more
expensive than usual macroscopic models due to the additional energy variable. In this
section we derive a macroscopic energy-transport model from the SHE model following [8].
We start with the scaled SHE equations of Theorem 4.14
N()
t
F divJ V
J

= S(F), x R
3
, R, t > 0, (4.87)
J(x, , t) = D(x, )
_
F +V
F

_
, (4.88)
for the distribution function F(x, , t) depending on the position x, energy and time t.
We have assumed that the eective eld in Theorem 4.14 can be written as E
e
= V
with the electrostatic potential V (x, t) which is assumed to be a given function. The
density of states N(), the diusion matrix D(x, ), and the averaged inelastic collision
operator S(F) are dened in (4.73), (4.74), and (4.75), respectively. We recall that the
inelastic collision operator contains the inelastic corrections to phonon scattering and
electron-electron collisions.
103
Our main assumption is that the electron-electron collision operator is dominant in
the sense
S(F) =
1

S
ee
(F) + S
ph
(F),
where S
ee
and S
ph
(F) are given by
(S
ee
(F))(x, , t) =
_
B
(Q
ee
(F))(x, k, t)( E(k)) dk,
(S
ph
(F))(x, , t) =
_
B
(Q
ac
(F) +Q
op
(F))(x, k, t)( E(k)) dk,
and the collision operators Q
ee
, Q
ac
and Q
op
model electron-electron scattering and scat-
tering with acoustic or optical phonons and are dened by (3.30) and (3.29), respectively.
Our assumption means that the energy loss due to phonon scattering occurs on a longer
time scale than carrier-carrier collisions. Note that the rescaled SHE equation
N()
t
F divJ V
J

=
1

S
ee
(F) + S
ph
(F) (4.89)
corresponds to the hyperbolic scaling of Section 4.2 (see (4.33)).
We need the following properties of the operators
(S
ee
(F))(x, , t) =
_
B
4
s
ee
(k, k
t
, k
1
, k
t
1
)[F
t
F
t
1
(1 F)(1 F
1
) FF
1
(1 F
t
)(1 F
t
1
)]
( E(k)) dk dk
t
dk
1
dk
t
1
, (4.90)
(S
ph
(F))(x, , t) =
_
B
2
[s
ph
(k
t
, k)F
t
(1 F) s
ph
(k, k
t
)F(1 F
t
)]( E(k)) dk dk
t
,
where F = F(E(k)), F
t
= F(E(k
t
)), F
1
= F(E(k
1
)), F
t
1
= F(E(k
t
1
)), and
s
ee
(k, k
t
, k
1
, k
t
1
) =
ee
(x, k, k
t
, k
1
, k
t
1
)(E(k) + E(k
1
) E(k
t
) E(k
t
1
))
has the properties
s
ee
(k, k
t
, k
1
, k
t
1
) = s
ee
(k
1
, k
t
, k, k
t
1
) = s
ee
(k
t
, k, k
t
1
, k
1
). (4.91)
Lemma 4.19 Let S
ee
(F) and S
ph
(F) be given as above satisfying (4.91). Then
_
R
S
ee
(F)
i
d = 0 (i = 0, 1),
_
R
S
ph
(F)d = 0.
Proof: The second assertion follows directly by integrating over R and using the
denition of the delta distribution:
_
R
S
ph
(F)d =
_
B
2
[s
ph
(k
t
, k)F
t
(1 F) s
ph
(k, k
t
)F(1 F
t
)]) dk dk
t
=
_
B
2
s
ph
(k
t
, k)F
t
(1 F) dk dk
t

_
B
2
s
ph
(k
t
, k)F
t
(1 F) dk
t
dk
= 0.
104
To prove the rst assertion we show a more general result. It holds for all F and G:
_
R
S
ee
(F)Gd =
1
4
_
B
4
s(k, k
t
, k
1
, k
t
1
)[F
t
F
t
1
(1 F)(1 F
1
) FF
1
(1 F
t
)(1 F
t
1
)]
(G
t
+G
t
1
GG
1
) d
4
k, (4.92)
where d
4
k = dk dk
t
dk
1
dk
t
1
. Indeed, we obtain from (4.91)
_
R
S
ee
(F)Gd
=
_
B
4
Q
ee
(F(E(k))G(E(k)) dk
=
1
4
_
B
4
s
ee
(k, k
t
, k
1
, k
t
1
)[F
t
F
t
1
(1 F)(1 F
1
) FF
1
(1 F
t
)(1 F
t
1
)]Gd
4
k
+
1
4
_
B
4
s
ee
(k
t
, k, k
t
1
, k
1
)[FF
1
(1 F
t
)(1 F
t
1
) F
t
F
t
1
(1 F)(1 F
1
)]G
t
d
4
k
+
1
4
_
B
4
s
ee
(k
1
, k
t
, k, k
t
1
)[F
t
F
t
1
(1 F
1
)(1 F) F
1
F(1 F
t
)(1 F
t
1
)]G
1
d
4
k
+
1
4
_
B
4
s
ee
(k
t
1
, k
1
, k
t
, k)[F
1
F(1 F
t
1
)(1 F
t
) F
t
1
F
t
(1 F
1
)(1 F)]G
t
1
d
4
k
=
1
4
_
B
4
s
ee
(k, k
t
, k
1
, k
t
1
)[F
t
F
t
1
(1 F)(1 F
1
) FF
1
(1 F
t
)(1 F
t
1
)]
(GG
t
+ G
1
G
t
1
) d
4
k.

It remains to show that the rst assertion follows from (4.92). Taking G() = 1 it is
clear that
_
R
S
ee
(F) d = 0.
Choosing G() = , we obtain an integral over B
4
involving the product
(E +E
1
E
t
E
t
1
)(E
t
+ E
t
1
E E
1
)
which vanishes in view of the properties of the distribution. This shows that
_
R
S
ee
(F) d = 0.
Lemma 4.20 (1) The kernel of S
ee
is given by
N(S
ee
) = F : R R : R, T > 0 : F = F
,T
,
where
F
,T
() =
1
1 + e
()/T
.
105
(2) For given F
0
, the equation DS
ee
(F
0
)F = G is solvable if and only if
_
R
G() d = 0 and
_
R
G() d = 0,
where DS
ee
(F
0
) is the rst derivative of S
ee
with respect to F
0
.
The variable is called chemical potential and T the electron temperature. The function
F
,T
is termed Fermi-Dirac distribution and has already been derived in Theorem 3.7
under dierent assumptions on the collision operator.
Proof: We only give a sketch of the proof and refer to [8, Sec. 4.2] for details.
(1) Choosing G = H(F) = ln F ln(1 F) in (4.92) yields
_
R
S
ee
(F)H(F) d
=
1
4
_
B
4
s
ee
(k, k
t
, k
1
, k
t
1
)[F
t
F
t
1
(1 F)(1 F
1
) FF
1
(1 F
t
)(1 F
t
1
)]
[ln(F
t
F
t
1
(1 F)(1 F
1
)) ln(FF
1
(1 F
t
)(1 F
t
1
))] d
4
k
0
since (x y)(ln x ln y) 0 for all x, y 0. Thus, taking F N(S
ee
), we obtain
ln(F
t
F
t
1
(1 F)(1 F
1
)) = ln(FF
1
(1 F
t
)(1 F
t
1
))
or
H(F
t
) + H(F
t
1
) = H(F) +H(F
1
)
when E(k
t
) +E(k
t
1
) = E(k) +E(k
1
) (which follows from the distribution in (4.90)) and
k
t
+ k
t
1
= k + k
1
modulo B (which follows from the so-called umklapp processes, see [8,
(2.5)]). It can be motivated that
H(F( c)) + H(F( + c)) = H(F()) + H(F())
for some c > 0. This means that the function x g(x) := H(F(x)) fullls the functional
equation
g(x c) + g(x +c) = 2g(x) for all x R and some c = c(x).
The only functions solving such an identiy are ane, i.e. g(x) = ax+b. Then, introducing
formally
T = 1/a and = b/a,
we obtain from
ln
F
1 F
= H(F) = a + b =

T
106
the equation
F() =
1
1 + e
()/T
.
(2) The assertion is a consequence of the Fredholm alternative Lemma 4.3; we refer
to [8, Prop. 4.2].
We now employ the Hilbert expansion
F = F
0
+ F
1
+
2
F
2
+ , J = J
0
+J
1
+
2
J
2
+ (4.93)
and Taylor approximation for the collision operators in (4.89):
N()
t
F
0
divJ
0
V
J
0

=
1

(S
ee
(F
0
) +DS
ee
(F
0
)F
1
)
+S
ph
(F
0
) + DS
ph
(F
0
)F
1
+ O(
2
).
Identifying equal powers of leads to
S
ee
(F
0
) = 0, N() =
t
F
0
divJ
0
V
J
0

= DS
ee
(F
0
)F
1
+ S
ph
(F
0
). (4.94)
Theorem 4.21 The Hilbert expansion (4.93) is solvable up to rst order if and only if
F
0
(x, , t) = F
(x,t),T(x,t)
() and (x, t), T(x, t) satisfy formally the equations

t
n(, t) divJ
n
= 0,

t
(ne)(, t) divJ
e
+V J
n
= W(, T),
J
n
= D
11

T
_
+ D
12

_
1
T
_
D
11
V
T
,
J
e
= D
21

T
_
+ D
22

_
1
T
_
D
21
V
T
,
where
n(, T) =
_
R
F
,T
()N() d, (4.95)
ne(, T) =
_
R
F
,T
()N() d, (4.96)
are the electron density and energy density, respectively,
W(, T) =
_
R
S
ph
(F
,T
) d
is the relaxation term, and
D
ij
(x, , T) =
_
R
D(x, )F
,T
(1 F
,T
)
i+j2
d, i, j = 1, 2, (4.97)
are diusion matrices in R
33
.
The variables J
n
and J
e
are called the macroscopic partial and energy current densities,
respectively, and /T and 1/T are the so-called entropy variables.
107
Proof: From Lemma 4.20 (1) follows that the solutions of the rst equation in (4.94) are
given by
F
0
(x, , t) = F
,T
(),
where and T also depend on the parameters x and t. Lemma 4.20 (2) implies that the
second equation in (4.94) is solvable if and only if
_
R
_
N()
t
F
,T
divJ
0
V
J
0

S
ph
(F
,T
)
_

i
d = 0, i = 0, 1. (4.98)
Dening the macroscopic particle and energy current densities by
J
n
(x, t) =
_
R
J
0
(x, , t) d, J
e
(x, t) =
_
R
J
0
(x, , t) d,
we can reformulate (4.98) as

t
n divJ
n
= V
_
R
J
0

d +
_
R
S
ph
(F
,T
) d = 0,

t
(ne) divJ
e
= V
_
R
J
0

d +
_
R
S
ph
(F
,T
) d
= V J + W(, T),
using Lemma 4.19.
It remains to compute the uxes J
n
and J
e
. We interpret F
,T
() as a function of the
entropy variables u
1
= /T, u
2
= 1/T, and ,
F(u
1
, u
2
, ) = F
,T
().
Then
F(u
1
, u
2
, ) =
1
1 + e
(u
1
+u
2
)
and
F
u
1
=
e
(u
1
+u
2
)
(1 +e
(u
1
+u
2
)
)
2
= F
,T
(1 F
,T
),
F
u
2
=
e
(u
1
+u
2
)
(1 +e
(u
1
+u
2
)
)
2
= F
,T
(1 F
,T
),
F

=
u
2
e
(u
1
+u
2
)
(1 +e
(u
1
+u
2
)
)
2
=
1
T
F
,T
(1 F
,T
).
This yields
J
0
= D(x, )
_
F
,T
+V
F
,T

_
= D(x, )
_
F
u
1
u
1
+
F
u
2
u
2
+V
F

_
108
and
J
n
= D
11
u
1
+ D
12
u
2
D
11
V
T
,
J
e
= D
21
u
2
+ D
22
u
2
D
21
V
T
and completes the proof.
The matrix T = (D
ij
) R
66
has the following properties.
Proposition 4.22 It holds:
(1) The matrix T = (D
ij
) is symmetric. Moreover, D
12
= D
21
and D

ij
= D
ji
for
i, j = 1, 2.
(2) If the functions

k
E, E
k
E (4.99)
are linearly independent then T(x, , T) is symmetric, positive denite for any R
and T > 0.
The hypothesis (4.99) is a geometric assumption on the band structure. It requires
that the energy band has a real three-dimensional structure, excluding bands depending
only on one or two variables, for instance.
Proof: Part (1) follows from the denition (4.97) of the (33)-matrices D
ij
and Lemma 4.15.
For the proof of part (2) we choose z = (, ) R
6
with , R
3
such that ,= 0. Then,
by the symmetry of D(x, ) and the property (4.80), we compute
z

Tz =
_
R
z

_
D(x, ) D(x, )
D(x, )
2
D(x, )
_
zF
,T
(1 F
,T
) d
=
_
R
( + )

D(x, )( + )F
,T
(1 F
,T
) d

_
R
K
N()
_
B
3

i,j=1
(
i
+
i
)
E
k
i
E
k
j
(
j
+
j
)( E(k))F
,T
(1 F
,T
) dk d
=
_
B
K
N(E(k))
_
3

j=1
(
j
+ E(k)
j
)
E
k
j
_
2
F
,T
(1 F
,T
) dk
=
_
B
K
N(E(k))

_

k
E
E
k
E
_
z

2
F
,T
(1 F
,T
) dk.
109
The last integral is positive since otherwise,
_

k
E
E
k
E
_
z = 0, z ,= 0,
would imply that (
k
E, E
k
E) is linearly dependent which is excluded.
The following result shows that the name relaxation term for W(, T) is justied.
In fact, the temperature of the particles is expected to relax to the constant lattice tem-
perature T
L
= 1 if there are no other forces.
Proposition 4.23 The relaxation term W(, T) with phonon collision operator (4.75)
and (3.29) is monotone in the sense of operator theory, i.e.
W(, T)(T 1) 0.
Proof: We recall the denition of W(, T),
W(, T) =
_
R
S
ph
() d
=
_
R
_
B
2
[s
ph
(k
t
, k)F
t
(1 F) s
ph
(k, k
t
)F(1 F
t
)]( E(k)) dk dk
t
d,
where F = F
,T
,
s
ph
(k, k
t
) =
ph
(x, k, k
t
)[(1 +N
ph
)(E
t
E + E
ph
) + N
ph
(E
t
E + E
ph
)], (4.100)
and E = E(k), E
t
= E(k
t
), N
ph
= (e
E
ph
1)
1
, and E
ph
> 0 is the phonon energy. Since
F = (1 F)M, M = e
()/T
,
we obtain
W(, T) =
_
B
2
(1 F
t
)(1 F)[s
ph
(k
t
, k)M
t
s
ph
(k, k
t
)M]E(k) dk dk
t
.
Denition (4.100) and the properties (x) = (x) and (x x
0
)x = (x x
0
)x
0
yield
[s
ph
(k
t
, k)M
t
s
ph
(k, k
t
)M]E(k)
=
ph
(1 +N
ph
)(E E
t
+ E
ph
)M
t
E +N
ph
(E E
t
E
ph
)M
t
E

ph
(1 +N
ph
)(E E
t
E
ph
)ME N
ph
(E E
t
+ E
ph
)ME
=
ph
(E E
t
+E
ph
)[(1 +N
ph
)M
t
N
ph
M]E
+
ph
(E E
t
E
ph
)[N
ph
M
t
(1 +N
ph
)M](E
t
+ E
ph
).
110
Thus, exchanging k and k
t
in the second integral and using (x) = (x) again, some
terms cancel and we end up with
W(, T) =
_
B
2
(1 F
t
)(1 F)(E E
t
+ E
ph
)E
ph
[N
ph
M (1 +N
ph
)M
t
] dk dk
t
=
_
B
2
(1 F
t
)(1 F)(E E
t
+ E
ph
)E
ph
N
ph
[M e
E
ph
M
t
] dk dk
t
,
since 1 +N
ph
= e
E
ph
N
ph
. The distribution allows to substitute E in M to E
t
E
ph
such
that
M e
E
ph
M
t
= e
(E)/T
e
E
ph
e
(E

)/T
= e
(E

)/T
_
e
E
ph
/T
e
E
ph
_
for all E and E
t
satisfying E = E
t
E
ph
. We conclude that
W(, T)(T 1) =
_
B
2
(1 F
t
)(1 F)(E E
t
+ E
ph
)E
ph
N
ph
M
t

_
e
E
ph
/T
e
E
ph
_
(T 1) dk dk
t
0,
since the function x e
E
ph
/x
is decreasing and hence,
_
e
E
ph
/T
e
E
ph
_
(T 1) 0 for all T > 0.
More explicit expressions for the particle and energy densities and for the diusion
matrix can be derived for spherically symmetric energy bands and Maxwell-Boltzmann
statistics.
Example 4.24 (Spherically symmetric energy bands)
We impose the following simplifying hypotheses:
The distribution function F
,T
is approximated by e
()/T
. This is possible if
T.
The scattering rate of the elastic collision operator (4.59) depends only on x and
E(k).
The energy band is spherically symmetric and strictly monotone in [k[.
From Example 4.16 and denitions (4.95)(4.96) of the particle and energy densities
follows
n(, T) =
_
R
e
()/T
N() d = 2e
/T
_
R
_
()
t
()e
/T
d,
ne(, T) =
_
R
e
()/T
N() d = 2e
/T
_
R
_
()
t
()e
/T
d,
111
where we have used the expression (4.81) for N() and [k[
2
= (E(k)). From (4.82) the
diusion matrices become
D
ij
(, T) =
4
3
e
/T
_
R
()
i+j2
s(x, )
t
()
2
e
/T
d I,
where I R
33
is the identity matrix.
Example 4.25 (Parabolic band approximation)
We impose the same assumptions as in the previous example and, additionally,
The energy band is parabolic: E(k) = [k[
2
/2, k R
3
.
The scattering rate is given by the so-called Chen model [21, 24]: s(x, ) = s
0
(x)

.
The rst assumption implies that () = 2, 0. The second hypothesis means that
the relaxation time (4.85) in the formulation (4.83) of the elastic collision operator,
(x, ) =
1
s(x, )N()
=
1
s
0
(x)

N()
,
holds with = 1/2. Recall that N() = 4

2 by Example 4.17. Using


_

0
e
z
2
/2
z
2
dz =
1
2
_
R
e
z
2
/2
z
2
dz =
1
2
_
R
e
z
2
/2
dz =
_

2
gives
n(, T) = 4

2e
/T
_

0
e
/T

d = 4

2e
/T
T
3/2

2
_

0
e
z
2
/2
z
2
dz
= (2T)
3/2
e
/T
. (4.101)
Furthermore, since
_

0
e
z
2
/2
z
4
dz =
1
2
_
R
e
z
2
/2
z
4
dz =
3
2

2
(see (4.44)), we obtain
(ne)(, T) = 4

2e
/T
_

0
e
/T

3/2
d = 4

2e
/T
T
5/2
2
3/2
_

0
e
z
2
/2
z
4
dz
=
3
2
(2)
3/2
T
5/2
e
/T
=
3
2
nT.
The diusion matrices become (see Example 4.24)
D
ij
(x, , T) =
2
3
e
/T
s
0
(x)
_
R

i+j3/2
e
/T
d
=
2
3
e
/T
s
0
(x)
_
T
2
_
i+j3/2
T
_

0
z
2(i+j1)
e
z
2
/2
dz.
112
Therefore, since
_

0
e
z
2
/2
z
6
dz =
1
2
_
R
e
z
2
/2
z
6
dz =
5 3
2

2
(see (4.44)), we infer that
D(x, , T) =

3s
0
(x)
T
3/2
e
/T
_
I
3
2
TI
3
2
TI
15
4
T
2
I
_
=
n(, T)
6

2s
0
(x)
_
I
3
2
TI
3
2
TI
15
4
T
2
I
_
R
66
.
Notice that the matrix D(x, , T) can be identied by the (2 2)-matrix
D(x, , T) =
n(, T)
6

2s
0
(x)
_
1
3
2
T
3
2
T
15
4
T
2
_
.
Example 4.26 (Fokker-Planck relaxation term)
The relaxation term W(, T) can be written explicitely in terms of and T using the
Fokker-Planck approximation (4.86) and the assumptions of the previous example. Then,
by integrating by parts,
W(, T) =
_

0

_
s
1

1/2
N()
2
_
F
,T
+
F
,T

__
d
= 2(4)
2
s
1
_

0

3/2
e
()/T
_
1
1
T
__
d
= 2(4)
2
s
1
_

0

3/2
e
()/T
_
1
1
T
_
d
= 2(4)
2
s
1

3
4

T
5/2
e
/T
_
1
1
T
_
=
3
2
n(, T)(1 T)

0
,
where

0
=
1
4

2s
1
is called the energy relaxation time.
Remark 4.27 The energy-transport model of Theorem 4.21 ist written with physical
113
parameters as

t
n(, T)
1
q
divJ
n
= 0,

t
(ne)(, T)
1
q
divJ
e
+V J
n
= W(, T),
J
n
= D
11

_
q
k
B
T
_
+ D
12

_
1
k
B
T
_
D
11
qV
k
B
T
,
J
e
= D
21

_
q
k
B
T
_
+ D
22

_
1
k
B
T
_
D
21
qV
k
B
T
.
In the situation of Examples 4.25 and 4.26 we call this system of equations the Chen
model. In particular,
n(, T) =
_
2m

e
k
B
T

2
_
3/2
e
q/k
B
T
,
(ne)(, T) =
3
2
n(, T) k
B
T,
D(x, , T) =
q
3
6

2S
0
(x)
n(, T)
_
1
3
2
k
B
T
3
2
k
B
T
15
4
(k
B
T)
2
_
,
W(, T) =
3
2
n(, T)k
B
(T T
L
)

0
,
where T
L
is the lattice temperature and
0
= (
2
/2m

e
)
3/2
/2S
1
.
4.5 Relaxation-time limits
In this chapter we have derived four uid-type models: the drift-diusion, hydrodynamic,
SHE, and energy-transport equations (see Figure 4.1 for a summary). In this section we
show that the drift-diusion and energy-transport equations can be formally derived from
the hydrodynamic model by performing so-called relaxation-time limits.
We recall the unscaled hydrodynamic model with electrostatic potential:

t
n
1
q
divJ = 0 (4.102)

t
J
1
q
div
_
J J
n
_

qk
B
m

e
(nT) +
q
2
m

e
nV =
J

p
, (4.103)

t
(ne)
1
q
div[J(e + k
B
T)] + J V div
_

p
m

0
nk
B
T(k
B
T)
_
(4.104)
=
n

W
_
e
d
2
k
B
T
L
_
, x R
d
; t > 0,
where the energy density is
ne =
m

e
2q
2
[J[
2
n
+
d
2
nk
B
T.
114
Drift-Diusion

Energy-transport

Boltzmann
.
Hydrodynamic
.
SHE

energy relaxation- time limit


dominant
collisions
expansion)
electron-electron
(Hilbert
momentum
relaxation-time
limit
dominant
electron-electron
collision (Hilbert
expansion)
moment method
dominant elastic
collisions (Hilbert
expansion)
Figure 4.1: Hierarchy of classical semiconductor models. The drift-diusion model can be
also directly derived from the Boltzmann equation via a moment method (see [9]).
These equations deviate from (4.102)(4.104) by the inclusion of the heat conduction
term with the heat conductivity
0
(see the end of Section 4.2) and by the introduction of
two dierent relaxation times; the momentum relaxation time
p
and the energy relaxation
time
W
. The reason is that we wish to consider two dierent time scales for momentum
and energy relaxation.
In order to scale (4.102)(4.104) we choose the following reference values:
reference length L,
reference particle density N,
reference temperature T
L
(lattice temperature),
reference time
0
,
reference potential U
T
= k
B
T
L
/q,
reference particle current density J
0
= qNL/
0
.
We can choose, for instance, L as the diameter of the semiconductor domain and N
as the maximal value of the doping concentration. The reference time
0
is given by the
assumption that the thermal energy is of the same order as the geometric average of the
kinetic energies to cross the domain in time
0
,
p
, respectively:
k
B
T
0
=

e
_
L

0
_
2

e
_
L

p
_
2
.
115
With these reference values we dene the scaled variables:
x = Lx
s
, t =
0
t
s
, n = Nn
s
,
T = T
L
T
s
, V = U
T
V
S
, J = J
0
J
s
.
Replacing the dimensional variables in (4.102)(4.104) by the scaled ones and omitting
the index s we obtain

t
n divJ = 0, (4.105)

t
J div
_
J J
n
_
(nT) + nV = J, (4.106)

t
(ne) div[J(e + T)] + J V div (
0
nTT) =
n

_
e
d
2
_
, (4.107)
where
=

p

0
, =

W

0
,
and the scaled energy is given by
e =
[J[
2
2n
2
+
d
2
T. (4.108)
We consider the following limits:
(1) 0 and 0,
(2) 0 and xed,
(3) xed and 0.
(1) The combined limit 0 and 0 corresponds to the physical situation
when the kinetic energy needed to cross the domain in time
0
is assumed to be much
smaller than the thermal energy, and when the momentum and energy relaxation times
are assumed to be of the same order. This means
1
m

e
(L/
0
)
2
k
B
T
0
=

p

0
= and O(1) =

p

W
=

. (4.109)
The second condition implies that also 1. Equations (4.105)(4.107) become in the
limit 0 and 0

t
n divJ = 0, J = (nT) nV, e =
d
2
. (4.110)
Moreover, by (4.108), e = (d/2)T, and the temperature is T = 1. Then the limit equations
(4.110) are the drift-diusion equations.
116
(2) The limit 0 corresponds to the physical situation expressed by the rst
equation in (4.109). Then (4.105)(4.107) yield in the limit 0

t
n divJ = 0, J = (nT) nV (4.111)
and

t
_
d
2
nT
_
div
_
d + 2
2
JT +
0
nTT
_
+ J V =
d
2
n

(T 1). (4.112)
In the entropic variables u
1
= /T and u
2
= 1/T we can write in view of the expression
n = (2T)
d/2
e
/T
(see (4.95) for the case d = 3),
J =
(nT)
u
1
u
1
+
(nT)
u
2
u
2
nV = n
_

T
_
+
d + 2
2
nT
2

1
T
_
nV,
and
d + 2
2
JT +
0
nTT
=
d + 2
2
T
_
n
_

T
_
+
d + 2
2
nT
2

1
T
_
nV
_
+
0
nT
3

1
T
_
=
d + 2
2
nT
_

T
_
+
_
_
d + 2
2
_
2
+
0
_
nT
3

1
T
_

d + 2
2
nTV.
Thus we can write (4.111)(4.112) as the energy-transport model in the entropic variables
with the diusion matrix
D = nT
_
1
d+2
2
T
d+2
2
T
_
_
d+2
2
_
2
+
0
_
T
2
_
.
The determinant of this matrix is positive (for n, T > 0) if and only if
0
> 0. This shows
that the heat conduction term is necessary to obtain wellposedness of the energy-transport
equations. We recall that this term is introduced heuristically, and no justication has
been given.
(3) The formal limit 0 only gives the equation e = d/2 which means that the sum
of kinetic and thermal energy is constant in space and time. More interesting is the limit
0 in the energy-transport equations (4.111)(4.112). This yields formally T = 1 and

t
n divJ = 0, J = n nV,
which are the drift-diusion euqations.
The three relaxation-time limits are summarized in Figure 4.2.
117
0
0
0, 0
Energy-transport Drift-diusion
Hydrodynamic
Figure 4.2: Relaxation-time limits in the hydrodynamic and energy-transport model.
118
5 Quantum Kinetic Models
5.1 The Schrodinger equation
As the dimensions of a semiconductor device decrease, quantum mechanical eects have
to be included in the modeling of the transport phenomena. It is of great importance to
devise models which, on the one hand, are capable of describing quantum eects and,
on the other hand, are suciently simple to allow for ecient numerical simulations. We
start with the many-particle Schrodinger equation and follow [49, Sec. 1.4].
The motion of an electron ensemble consisting of M particles in a vacuum under the
inuence of a (real-valued) electrostatic potential V is described by the many-particle
Schrodinger equation,
i
t
=

2
2m
M

j=1

x
j
qV (x, t), x R
dM
, t > 0, (5.1)
(x, 0) =
I
(x), (5.2)
where x = (x
1
, . . . , x
M
)

R
dM
, i
2
= 1, = h/2 is the reduced Planck constant, m
the electron mass, q the elementary charge, and = (x, t) is called the wave function of
the electron ensemble. The Laplace operator
x
j
only acts on the position variable x
j
,

x
j
=
d

=1

x
j,
, where x
j
= (x
j,1
, . . . , x
j,d
)

R
d
.
The variable x
j
R
d
denotes the position vector of the j-th electron of the ensemble. The
ensemble position density n and the ensemble current density J are dened by
n = [[
2
, J =
q
m
Im(
x
),
where is the complex conjugate of .
There are two alternative formulations of the motion of the electron ensemble:
the density-matrix formulation and
the kinetic (Wigner) formulation.
We introduce the density matrix in this section and explain the kinetic formulation in the
following section.
Dene the ensemble Hamilton operator
H = H
x
=

2
2m

x
qV (x, t) =

2
2m
M

j=1

x
j
qV (x, t)
119
and the density matrix
(r, s, t) = (r, t)(s, t), r, s R
dM
, t > 0. (5.3)
Then the ensemble densities can be written in terms of as
n(x, t) = (x, x, t), (5.4)
J(x, t) =
iq
2m
(
x

x
)(x, t) =
iq
2m
(
s

r
)(x, x, t). (5.5)
The evolution equation for is obtained by dierentiating (5.3) and using the Schrodinger
equation
i
t
= H (5.6)
which gives
i
t
(r, s, t) = i(
t
(r, t)(s, t) + (r, t)
t
(s, t))
= (H
s
H
r
)(r, s, t), (5.7)
since V (x, t) is real-valued. This equation is referred to as the Liouville-von Neumann or
Heisenberg equation. With the initial condition
(r, s, 0) =
I
(r)
I
(s), r, s R
dM
, (5.8)
the Schrodinger problem (5.1)(5.2) and the Heisenberg Problem (5.7)(5.8) are formally
equivalent.
Is there any relation between the Schrodinger and the Heisenberg picture in the case
of general initial conditions
(r, s, 0) =
I
(r, s), r, s R
dM
? (5.9)
In order to derive a relation let (
j
) be an orthonormal basis of L
2
(R
dM
). Then (
j

k
) is
an orthonormal basis of L
2
(R
dM
R
dM
) and we can develop

I
(r, s) =

j,k=1

jk

j
(r)
k
(s),
if
I
L
2
(R
dM
R
dM
) is assumed. Now let
j
be the solution of the Schrodinger equation
(5.6) with initial datum
j
. A computation as above shows that the product
j
(r, t)
k
(s, t)
solves the Heisenberg equation (5.7), and since this equation is linear, also
(r, s, t) =

j,k=1

jk

j
(r, t)
k
(s, t) (5.10)
120
solves the Heisenberg equation with initial datum
I
. Thus, the solution of the Heisenberg
problem (5.7), (5.9) can be written as an innite sum of wave functions. Such a situation
is called mixed quantum states, whereas the formulation (5.3) refers to a single quantum
state.
In the mixed state case (5.10), the ensemble position density is written as
n(x, t) =

j=1

jj
[
j
(x, t)[
2
,
and the weights
jj
can be interpreted as occupation probabilites of the j-th quantum
state.
5.2 The quantum Liouville equation
The Wigner equation is the quantum equivalent of the Liouville equation and can be de-
rived from the Heisenberg equation by Fourier transformation. We recall that the Fourier
transform F is dened by
g() = (Fg)() = (2)
dM/2
_
R
dM
g(v)e
iv
dv
for (suciently smooth) functions g : R
dM
C with inverse

h(v) = (F
1
h)(v) = (2)
dM/2
_
R
dM
h()e
iv
d
for functions h : R
dM
C.
Denition 5.1 The inverse Fourier transform of the density matrix,
w(x, v, t) = (2)
dM
_
R
dM

_
x +

2m
, x

2m
, t
_
e
iv
d, (5.11)
is called Wigner function.
The Wigner function has been introduced by Wigner in 1932 [66]. Setting
u(x, , t) =
_
x +

2m
, x

2m
, t
_
, (5.12)
the expression
w = (2)
dM/2
u
shows that w is indeed related to the inverse Fourier transform of . Since (/2m) has
the dimension of length and /2m has the dimension of m
2
/s, has the unit of inverse
velocity s/m. Thus, the variable v in the Fourier transform has the dimension of velocity
m/s.
121
Lemma 5.2 Let be a solution of the Heisenberg equation (5.7), (5.9). Then the Wigner
function w, dened in (5.11), is formally a solution of

t
w + v
x
w
q
m
[V ]w = 0, x, v R
dM
, t > 0, (5.13)
w(x, v, 0) = w
I
(x, v), x, v R
dM
,
where
w
I
(x, v) = (2)
dM
_
R
dM

I
_
x +

2m
, x

2m

_
e
iv
d
and [V ] is a so-called pseudo-dierential operator, dened by
(

[V ]w)(x, , t) = V (x, , t) w(x, , t), (5.14)


where
V (x, , t) =
im

_
V
_
x +

2m
, t
_
V
_
x

2m
, t
__
.
Equation (5.13) is called the many-particle Wigner equation or quantum Liouville
equation. Generally an operator whose Fourier transform acts as a multiplication operator
on the Fourier transform of its argument, is called a (linear) pseudo-dierential operator.
We refer to [62] for the mathematical theory of pseudo-dierential operators. The function
V is a kind of discrete directional derivative. Indeed, in the formal limit 0 it holds
V (x, , t) i
x
V (x, t) = i
V

(x, t). (5.15)


Clearly, the limit 0 makes sense only after an appropriate scaling (see below).
The operator [V ]w is written more explicitly as
([V ]w)(x, v, t) = ((V ) w)

(x, v, t) = (2)
dM/2
((V )u)

(x, v, t) (5.16)
=
1
(2)
dM
_
R
dM
V (x, , t)u(x, , t)e
iv
d
=
1
(2)
dM
_
R
dM
_
R
dM
V (x, , t)w(x, v
t
, t)e
i(vv

)
dv
t
d. (5.17)
Proof: Since, by (5.12),
div

(
x
u)(x, , t) = div

(
r
+
s
)
_
x +

2m
, x

2m
, t
_
=

2m
(
r

s
)
_
x +

2m
, x

2m
, t
_
,
122
the transformed Heisenberg equation for u becomes, for r = x +/2m, s = x /2m,
i
t
u(x, , t) = i
t
(r, s, t)
=
_

2
2m
(
s

r
) qV (s, t) +qV (r, t)
_
(r, s, t)
= div

(
x
u)(x, , t) +
iq
m
V (x, , t)u(x, , t)
or

t
u + idiv

(
x
u)
q
m
(V )u = 0, x, R
dM
, t > 0.
Taking the inverse Fourier transform gives

t
u + i(div

x
u)

q
m
((V )u)

= 0. (5.18)
We compute the second and the third term of this equation:
i(div

x
u)

(x, v, t) =
i
(2)
dM/2
_
R
dM
div

x
u(x, , t)e
iv
d
=
v
(2)
dM/2
_
R
dM

x
u(x, , t)e
iv
d
= v
x
u(x, v, t) = (2)
dM/2
v
x
w(x, v, t),
where we have employed integration by parts, and, by (5.16),
([V ]w)(x, v, t) = (2)
dM/2
((V )u)

(x, v, t).
Therefore, (5.18) equals the Wigner equation (5.13).
Lemma 5.3 The ensemble particle density n and the ensemble current density J can be
expressed in terms of the Wigner function as
n(x, t) =
_
R
dM
w(x, v, t) dv, J(x, t) = q
_
R
dM
w(x, v, t)v dv.
The above integrals are called the zeroth and rst moments of the Wigner function,
respectively, in analogy to the classical situation (see Section 3.1).
Proof: The rst identity follows from (5.4):
n(x, t) = (x, x, t) = u(x, 0, t) = (2)
dM/2
w(x, 0, t) =
_
R
dM
w(x, v, t) dv.
For the proof of the second identity we use the fact that the Fourier transform translates
dierential operators into a multiplication:
i

w(x, , t) =
1
(2)
dM/2
_
R
dM
w(x, v, t)ve
iv
dv = vw(x, , t).
123
Therefore, by (5.5),
J(x, t) =
iq
2m
(
s

r
)(x, x, t) = iq

u(x, 0, t)
= iq(2)
dM/2

w(x, 0, t) = q(2)
dM/2
vw(x, 0, t)
= q
_
R
dM
w(x, v, t)v dv.

We discuss two questions related to the quantum Liouville equation:


How can we formalize the classical limit 0 and which is the limit equation?
Are the solutions of the quantum Liouville equation nonnegative if this property
holds true initially?
The limit 0 can be formalized in an appropriate scaling. We choose the reference
length , the reference time , the reference velocity /, and the reference voltage k
B
T/q,
where k
B
is the Boltzmann constant and T the temperature. We assume that the reference
wave energy / is much smaller than the thermal and kinetic energies,
/
k
B
T
= and
/
m(/)
2
= with 1
X
(this xes for given and vice versa). Thus, introducing the scaling
x = x
s
, t = t
s
, v =

v
s
, V =
k
B
T
q
V
s
,
we obtain, after omitting the index s,

t
w + v
x
w [V ]w = 0, (5.19)
where [V ]w is given by (5.17) with
V (x, , t) =
i

_
V (x +

2
, t) V (x

2
, t)
_
.
The limit 0 leads to (see (5.15))
V (x, , t) i
x
V (x, t)
and hence, using i(u)

=
v
u,
[V ]w = (2)
dM/2
((V )u)

i(2)
dM/2
((
x
V )u)

= (2)
dM/2

x
V
v
u =
x
V
v
w.
124
Thus, (5.19) becomes in the limit 0

t
w + v
x
w
x
V
v
w = 0,
which is the classical Liouville equation (see Section 3.1). The above consideration makes
clear the limit performed in (5.15). The classical limit 0 has been made rigorous
in [47, 48] for smooth potentials.
The solution of the classical Liouville equation stays nonnegative for all time if the
initial distribution function is nonnegative. Unfortunately, this property does not hold for
the quantum Liouville equation. Thus, a (fully) probabilistic interpretation of the Wigner
function as a distribution function is not possible. In the case of a pure quantum state it
is possible to characterize the states for which the Wigner function is nonnegative exactly.
It is shown in [38] that
w(x, v, t) =
1
(2)
dM
_
R
dM

_
x +

2m
, t
_

_
x

2m
, t
_
e
iv
d
is nonnegative if and only if either 0 or
(x, t) = exp
_
x

A(t)x a(t) x b(t)


_
, x R
dM
, t > 0,
where A(t) C
dMdM
is a matrix with symmetric positive denite real part and a(t)
C
dM
, b(t) C. Inserting this ansatz into the Schrodinger equation (5.1) shows that the
potential has to be quadratic in x, i.e.
V (x, t) = x

A(t)x +a(t) x +

b(t)
for some

A(t) C
dMdM
, a(t) C
dM
,

b(t) C, in order to obtain a nonnegative Wigner


solution.
The case of mixed quantum states, i.e. for arbitrary initial data w
I
L
2
(R
dM
R
dM
),
is more involved. In fact, a necessary condition for the nonnegativity of w is not known.
5.3 The quantum Vlasov and Boltzmann equation
The quantum Liouville equation has the same disadvantage as its classical analogue,
namely, the equation needs to be solved in a very high-dimensional phase space which
makes its numerical solution almost unfeasible. In this section we derive the quantum
analogue of the classical Vlasov equation, the quantum Vlasov equation, which acts on a
2d-dimensional phase space. We proceed similarly as in [49, Sec. 1.5] (also see Section 3.2).
Consider as in the proceeding section an ensemble of M electrons with mass m in a
vacuum under the action of a (real-valued) electrostatic potential V (x, t). The motion of
the particle ensemble is described by the density matrix as a solution of the Heisenberg
equation (5.7). We impose the following assumptions:
125
(1) The potential can be decomposed into a sum of external potentials acting on one
particle and of two-particle interaction potentials,
V (x
1
, . . . , x
M
, t) =
M

j=1
V
ext
(x
j
, t) +
1
2
M

i,j=1
V
int
(x
i
, x
j
), (5.20)
where V
int
is symmetric, i.e. V
int
(x
i
, x
j
) = V
int
(x
j
, x
i
) for all i, j = 1, . . . , M, and is
of the order 1/M as M .
(2) The electrons of the ensemble are initially indistinguishable in the sense
(r
1
, . . . , r
M
, s
1
, . . . , s
M
, 0) = (r
(1)
, . . . , r
(M)
, s
(1)
, . . . , s
(M)
, 0) (5.21)
for all permutations of 1, . . . , M and all r
j
, s
j
R
d
.
(3) The initial subsensemble density matrices can be factorized

(a)
I
(r
(a)
, s
(a)
) =
(a)
(r
(a)
, s
(a)
, 0) =
a

j=1
R
I
(r
j
, s
j
), 1 a < M 1.
We discuss now these assumptions. The factor
1
2
in (5.20) is necessary since each
electron-electron pair in the sum of two-particle interactions is counted twice. The sym-
metry of the interaction potentials implies that
V (x
1
, . . . , x
M
, t) = V (x
(1)
, . . . , x
(M)
, t) for all t 0
and for all permutations . This property and (5.21) has the consequence that
(r
1
, . . . , r
M
, s
1
, . . . , s
M
, t) = (r
(1)
, . . . , r
(M)
, s
(1)
, . . . , s
(M)
, t) (5.22)
holds for all t > 0. Physically, this means that the electrons are indistinguishable for all
time.
In fact, (5.22) is not enough to describe the behavior of the electron ensemble. Indeed,
in the pure quantum state,
(r
1
, . . . , r
M
, s
1
, . . . , s
M
, t) = (r
1
, . . . , r
M
, t)(s
1
, . . . , r
M
, t),
where is the electron-ensemble wave function, the condition (5.22) is satised if either
the wave function is antisymmetric,
(x
1
, . . . , x
M
, t) = sign()(x
(1)
, . . . , x
(M)
, t)
for all permutations , where sign() 1, +1 is the sign of , or if it is symmetric,
(x
1
, . . . , x
M
, t) = (x
(1)
, . . . , x
(M)
, t)
126
for all . The rst case corresponds to ensembles consisting of Fermions (and in particular,
of electrons) and the latter case to ensembles of Bosons. In the rst case (which is the
interesting case for us) we obtain
(x
1
, . . . , x
M
, t) = 0 if x
i
= x
j
for some i ,= j.
This expresses the Pauli exclusion principle which is not contained in condition (5.22).
We wish to derive the evolution equation for a subensemble density matrix. The density
matrix of a subensemble consisting of a electrons is dened by

(a)
(r
(a)
, s
(a)
, t) =
_
R
d(Ma)
(r
(a)
, u
a+1
, . . . , u
M
, s
(a)
, u
a+1
, . . . , u
M
, t) du
a+1
. . . du
M
,
where
r
(a)
= (r
1
, . . . , r
a
), s
(a)
= (s
1
, . . . , s
a
) R
da
.
Clearly, in view of the indistinguishable property (5.22), the subensemble density matrices
satisfy

(a)
(r
1
, . . . , r
a
, s
1
, . . . , s
a
, t) =
(a)
(r
(1)
, . . . , r
(a)
, s
(1)
, . . . , s
(a)
, t) (5.23)
for all permutations of 1, . . . , a and all r
j
, s
j
R
d
, t 0.
We recall that the evolution of the complete electron ensemble is governed by the
Heisenberg equation (5.7), rewritten as
i
t
=

2
2m
M

j=1
(
s
j

r
j
) q
M

j=1
(V
ext
(s
j
, t) V
ext
(r
j
, t))

q
2
M

j,=1
(V
int
(s
j
, s

) V
int
(r
j
, r

)) . (5.24)
Theorem 5.4 Let the above hypotheses (1)(3) hold. Then the density matrix (r, s, t)
is a particular solution to the Heisenberg equation (5.23) if M 1,

(a)
(r
(a)
, s
(a)
, t) =
a

j=1
R(r
j
, s
j
, t), (5.25)
and the function W = (2)
d/2
M

R is a solution to the quantum Vlasov equation

t
W + v
x
W
q
m
[V
e
]W = 0, x, v R
d
, t > 0 (5.26)
W(x, v, 0) = W
I
(x, v), x, v R
d
,
where the pseudo-dierential operator [V
e
] is dened as in (5.14), the eective potential
V
e
is
V
e
(x, t) = V
ext
(x, t) +
_
R
d
n(z, t)V
int
(x, z) dz, (5.27)
127
the quantum electron density is
n(x, t) =
_
R
W(x, v, t) dv = MR(x, x, t), (5.28)
and W
I
= (2)
d/2
M

R
I
.
The expression (5.25) is also called Hartree ansatz. As the eective potential depends
on the function W through (5.28), we obtain a nonlinear pseudo-dierential equation.
Proof: We set u
j
= s
j
= r
j
for j = a +1, . . . , M in (5.24), integrate over (u
a+1
, . . . , u
M
)
R
d(Ma)
and use the property (5.23) to obtain, after an analogous computation as in
Section 3.2, a quantum equivalent of the BBGKY hierarchy,
i
t

(a)
=

2
2m
M

j=1
(
s
j

r
j
)
(a)
q
M

j=1
(V
ext
(s
j
, t) V
ext
(r
j
, t))
(a)
q(M a)
a

j=1
_
R
d
(V
int
(s
j
, u

) V
int
(r
j
, u

))
(a+1)

du

for all 1 a M 1, where

(a+1)

=
(a+1)
(r
(a)
, u

, s
(a)
, u

, t).
Since we have assumed that V
int
is of the order of 1/M as M , we can approximate
the above equation, by neglecting terms of order 1/M, by
i
t

(a)
=

2
2m
a

j=1
(
s
j

r
j
)
(a)
q
a

j=1
(V
ext
(s
j
, t) V
ext
(r
j
, t))
(a)
q
a

j=1
_
R
d
(V
int
(s
j
, u

) V
int
(r
j
, u

)) M
(a+1)

du

.
Similar to the classical case, it can be seen that this equation is satised by the ansatz
(5.25) if R solves the equation
i
t
R =

2
2m
(
s

r
)R q (V
e
(s, t) V
e
(r, t)) R, r, s, R
d
, t > 0, (5.29)
where the eective potential V
e
is dened in (5.27) and (5.28).
The kinetic formulation of (5.29) is derived as in Section 5.2. Introducing the change
of coordinates
r = x +

2m
, s = x

2m
,
the function U(x, , t) = MR(r, s, t) solves the equation

t
U + idiv

x
U +
iq

_
V
e
_
x +

2m
, t
_
V
e
_
x

2m
, t
__
U = 0.
128
Finally, the inverse Fourier transform W = (2)
d/2

U solves the quantum Vlasov equation
(5.26).
Contrary to the classical Vlasov equation, the quantum Vlasov equation does not
preserve the nonnegativity of the solution; cf. the discussion in Section 5.2. However, if
the initial single-particle density matrix R(r, s, 0) is positive semi-denite, the electron
density n, dened in (5.28), remains nonnegative for all times.
Similar to the quantum Liouville equation, the solution to the quantum Vlasov equa-
tion converges (at least formally) as 0 to a solution to the classical Vlasov equation

t
W +v
x
W
q
m

x
V
e

v
W = 0.
The limit 0 has to be understood in the sense explained in Section 5.2.
As in Section 3.2, a usual choice for the two-particle interaction potential is the
Coulomb potential which reads in three space dimensions as
V
int
(x, y) =
q
4
s
1
[x y[
, x, y R
3
, x ,= y,
where
s
denotes the permittivity of the semiconductor material (see (3.19)). In Section 3.2
it has been shown that the eective potential
V
e
(x, t) = V
ext
(x, t)
q
4
s
_
R
3
n(z, t)
[z x[
dz (5.30)
solves the Poisson equation

s
V
e
= q(n C(x)),
where C(x) = (
s
/q)V
ext
(x) is the doping concentration if V
ext
is generated by ions of
charge +q in the semiconductor. This concept can be generalized to any space dimen-
sion. The initial-value problem (5.26), (5.30) (together with (5.28)) is called the quantum
Vlasov-Poisson problem.
The presented quantum Vlasov equation models the motion of an electron ensemble
in a vacuum under the inuence of long-range interactions. However, the electrons are
moving in a crystal whose inuence needs to be taken into account. Including a lattice
potential into the Schrodinger equation, making a Bloch decomposition (see Theorem 2.5),
and letting the length scale of the Brillouin zone to innity, it can be shown [2, 55] that
the resulting equation equals

t
w + v(k)
x
w
q
m
[V
e
]w = 0, x, k R
d
, t > 0, (5.31)
together with equations (5.27) and (5.28) and initial conditions for w. Here, v(k) =

k
E(k)/ is the group velocity, E(k) the band structure, and k the wave vector which
has been extended to the whole space by the limiting procedure.
129
A second generalization of the quantum Vlasov equation is given by the inclusion
of short-range interactions modeled by scattering events of particles. The quantum me-
chanical modeling of collisions of electrons (with phonons, for instance) is a very dicult
task, and there is no complete theory up to now. The phenomenological approach used in
Section 3.3 cannot be used here since the notion of particle trajectories or characteristics
does not make sense in this quantum mechanical framework. On the other hand, most of
the collision models derived in the literature are highly complicated and numerically too
expensive [18, 31]. A simple approach is to formulate the so-called quantum Boltzmann
equation

t
w + v
x
w
q
m
[V
e
]w = Q(w), x, v R
d
, t > 0, (5.32)
by adding a heuristic collision term to the right-hande side of the quantum Vlasov equation
(5.26) or of its energy-band version (5.31). In numerical studies, often one of the following
collision operators are used (assuming the energy-band version of the quantum Boltzmann
equation)
the relaxation-time or BGK model [59]
Q(w) =
1

_
n
n
0
w
0
w
_
,
where
n(x, t) =
_
R
d
w(x, k, t) dk, n
0
(x, t) =
_
R
d
w
0
(x, k, t) dk,
and w
0
is the distribution function of the quantum mechanical thermal equilibrium,
dened, for instance in the mixed state, by the thermal equilibrium density matrix

0
(r, s) =

j
f(
j
)
j
(r)
j
(s),
where (
j
,
j
) are eigenvalue-eigenfunction pairs of the quantum Hamiltonian, and
f() is the Fermi-Dirac distribution;
the Caldeira-Leggett model [16]
Q(w) =
1

0
div
k
_
m

k
B
T

2

k
w +kw
_
, (5.33)
where
0
denotes the relaxation time, T the lattice temperature and m

the eective
electron mass.
The Caldeira-Leggett model has the disadvantage that the positivity of the density
matrix is not preserved under temporal evolution. In [17] the approach of Caldeira and
Leggett has been improved by deriving the Fokker-Planck operator
Q(w) =
1

0
div
k
_
m

k
B
T

2

k
w + kw
_
+
1

0
div
x
_

6k
B
T

k
w +

2
12m

k
B
T

x
w
_
,
130
where is the cut-o frequency modeling the interaction of the electrons with the phonons
of the crystal (see [17] for details). This collision operator preserves the posivitiy of the
density matrix.
A summary of the kinetic models derived in Chapter 3 and this chapter is presented
in Figure 5.1. Notice that for each model, there is an energy-band version in the (x, k, t)
variables, which reduces to a model in the (x, v, t) variables in the parabolic band approx-
imation via v = v(k) = k/m

.
quantum Boltzmann equation Boltzmann equation
quantum Vlasov equation Vlasov equation
quantum Liouville equation Liouville equation
no collisions no collisions
no two-particle interactions no two-particle interactions
0
0
0
Figure 5.1: Hierarchy of classical and quantum kinetic models.
5.4 Open quantum systems
Up to now, we have only considered closed quantum systems. In a closed system, all el-
ements and their interactions are completely known. On the other hand, open quantum
systems are characterized by the fact that elements of the system interact with an envi-
ronment but not all interactions are known precisely. As an example of an open system we
consider the one-dimensional stationary Schrodinger equation with some potential V (x),


2
2m

tt
qV (x) = E in (0, L), (5.34)
where m

is the eective mass which is assumed, for simplicity, to be constant. It is con-


venient to scale the Schrodinger equation. Choosing the reference length L, the reference
potential k
B
T/q, and the reference energy k
B
T and introducing the scaled variables
x = Lx
s
, V =
k
B
T
q
V
s
, E = (k
B
T)E
s
,
we obtain from (5.34), after omitting the index s,

2
2

tt
V (x) = E, x (0, 1), (5.35)
131
where
=

L
2
k
B
T
is the scaled Planck constant.
Equation (5.35) is an eigenvector-eigenvalue problem. For its solution we need to
specify boundary conditions. Since we do not know the wave function at the boundary or
outside of the interval, this constitutes an open system. In order to solve it, we have to
specify what happens with the electron wave outside of the interval. We assume: Electron
waves are injected at x = 0; they exit the interval at x = L or they are reected by the
potential at x = 0 (Figure 5.2). We solve the problem (5.35) by extending the potential
to the whole line,
V (x) = V (0) for x < 0, V (x) = V (1) for x > 1, (5.36)
and solve (5.35) for x R. Since the potential is constant in x < 0 and x > 1,
respectively, we expect that the solutions are plane waves in these intervals. This motivates
the following ansatz. Let rst p > 0. Then we dene

p
(x) =
_
e
ipx/
+ r(p)e
ipx/
: x < 0
t(p)e
ip
+
(p)(x1)/
: x > 1,
(5.37)
where p
+
(p) has to be determined. This ansatz means that a wave with amplitude one
is coming from (since p > 0) and is either transmitted to + with amplitude t(p)
or it is reected by the potential and travels back to with amplitude r(p). Inserting
(5.37) into the scaled Schrodinger equation (5.35) yields for x < 0
(E + V (0))
p
=

2
2

tt
p
=
p
2
2
e
ipx/
+
p
2
2
r(p)e
ipx/
=
p
2
2

p
and thus
E
p
= E =
p
2
2
V (0);
for x > 1 we obtain
(E
p
+ V (1))
p
=

2
2

tt
p
=
p
2
+
(p)
2

p
and
p
+
(p) =
_
2(E
p
+ V (1)) =
_
p
2
+ 2(V (1) V (0)).
We take the positive root since the wave travels to + and hence, p
+
(p) > 0 is required.
For the case p < 0 we make an analogous ansatz:

p
(x) =
_
t(p)e
ip

(p)x/
: x < 0
e
ip(x1)/
+ r(p)e
ip(x1)/
: x > 1,
132
injected
electrons electrons
electrons
reected
transmitted
0 L
x
p > 0
Figure 5.2: Electrons with p > 0 are injected at x = 0 and are reected at x = 0 or
transmitted at x = L.
where again p

(p) is unknown. This ansatz models a wave coming from + and being
either transmitted to or reected at x = 1 and traveling back to +. Inserting this
ansatz into (5.35) gives, after a similar computation as above,
E
p
= E =
p
2
2
V (1), p

(p) =
_
p
2
2(V (1) V (0)).
Assuming that the wave function
p
is continuous in R, this allows to specify the
boundary conditions for
p
at x = 0 and x = 1. Indeed, for p > 0 and x < 0, x 0 there
holds:

t
p
(0) = ip(1 r(p)) and ip
p
(0) = ip(1 +r(p)),
so

t
p
(0) +ip
p
(0) = 2ip.
At x > 1, x 1, we infer

t
p
(1) = ip
+
(p)t(p) = ip
+
(p)
p
(1).
For p < 0 we obtain

t
p
(1) ip
p
(1) = ip(1 r(p)) ip(1 +r(p)) = 2ip,

t
p
(0) = ip

(p)t(p) = ip

(p)
p
(0).
We have shown the following result.
Lemma 5.5 Let V (x), x (0, 1), be a given potential extended to R by (5.36). Then the
solution of the eigenvalue problem

2
2

tt
V (x) = E, x R,
133
can be written as
p > 0 : E
p
=
p
2
2
V (0),
p
(x) =
_
e
ipx/
+r(p)e
ipx/
: x 0
t(p)e
ip
+
(p)(x1)/
: x 1,
(5.38)
p < 0 : E
p
=
p
2
2
V (1),
p
(x) =
_
t(p)e
ip

(p)x/
: x 0
e
ip(x1)/
+r(p)e
ip(x1)/
: x 1,
(5.39)
where
p

(p) =
_
p
2
2(V (1) V (0)),
and in the interval (0, 1)
p
is the solution of

2
2

tt
p
V (x)
p
= E
p

p
, x (0, 1)

t
p
(0) +ip
p
(0) = 2ip,
t
p
(1) = ip
+
(p)
p
(1) for p > 0, (5.40)

t
p
(1) +ip
p
(1) = 2ip,
t
p
(0) = ip

(p)
p
(0) for p < 0, (5.41)
where E
p
is given as above. The reection and transmission amplitudes r(p) and t(p) are
determined by
p > 0 : r(p) =
1
2
_

p
(0) +
i
p

t
p
(0)
_
, t(p) =
p
(1), (5.42)
p < 0 : r(p) =
1
2
_

p
(1)
i
p

t
p
(1)
_
, t(p) =
p
(0). (5.43)
The boundary conditions (5.40)(5.41) are called the Lent-Kirkner boundary condi-
tions [41].
Proof: It remains to show the formulas (5.42)(5.43). In fact, they follow immediately
from (5.40) and (5.41). For instance, for p > 0, we have from the rst equation in (5.40)
i
p

t
p
(0) =
p
(0) 2
and hence, by (5.38),
1
2
_

p
(0) +
i
p

t
p
(0)
_
=
p
(0) 1 = r(p),
p
(1) = t(p).
In a similar way, (5.41) and (5.39) imply (5.43).
The unscaled macroscopic electron and current densities are given by
n(x) =
_
R
f(p)[
p
(x)[
2
dp
J(x) =
q
m

_
R
f(p)Im(
p
(x)
p
(x)) dp,
134
where f(p) describes the statistics of the mixed states. In Lemma 2.20 we have derived
the following formula:
f(p) =
m

k
B
T
2
2

3
ln
_
1 + e
(p
2
/2m

+E
F
)/k
B
T
_
,
where E
F
is the Fermi energy.
135
6 Quantum Fluid Models
6.1 Zero-temperature quantum hydrodynamic equations
We consider a single electron in a vacuum. The evolution of the particle is described by
the Schrodinger equation
i
t
=

2
2m
qV (x, t), x R
d
, t > 0,
(x, 0) =
I
(x), x R
d
. (6.1)
First we scale the equations by introducing reference values for the time, length, and
potential,
t = t
s
, x = Lx
s
, V = UV
s
.
We assume that the kinetic energy is of the same order as the electric energy,
m
_
L

_
2
= qU.
Then the scaled Schrodinger equation becomes (omitting the index s)
i
t
=

2
2
V , x R
d
, t > 0, (6.2)
where
=
/
m(L/)
2
=

mL
2
is the scaled Planck constant.
In order to derive a uid-type equation, we asume that the initial wave function is
given by the WKB (Wentzel, Kramers, Brillouin) state

I
=

n
I
exp(iS
I
/), (6.3)
where n
I
(x) 0, S
I
(x) R are some functions. We claim that the zero-temperature
quantum hydrodynamic equations

t
n divJ = 0, (6.4)

t
J div
_
J J
n
_
+ nV +

2
2
n
_

n
_
= 0, x R
d
, t > 0 (6.5)
n(x, 0) = n
I
(x), J(x, 0) = J
I
(x), x R
d
, (6.6)
are formally equivalent to the Schrodinger equation (6.2) in the sense of the following
theorem.
136
Theorem 6.1 (1) Let be a solution of (6.1)(6.2) with initial datum (6.3). Then
n = [[
2
, J = Im() solve (6.4)(6.6) with initial data
n
I
= [
I
[
2
, J
I
= n
I
S
I
(6.7)
as long as n > 0 in R
d
.
(2) Let (n, S) be a solution of

t
n+div(nS) = 0,
t
S+
1
2
[S[
2
V

2
2

n
= 0, x R
d
, t > 0, (6.8)
n(x, 0) = n
I
(x), S(x, 0) = S
I
(x), x R
d
,
such that n > 0 in R
d
, t > 0. then =

n exp(iS/) solves (6.1)(6.2) with initial


datum (6.3).
Notice that (6.8) implies (6.4)(6.5). Indeed, setting J = nS, we obtain imme-
diately (6.4), and dierentiation of the second equation in (6.8) with respect to x and
multiplication with n yields
nV +

2
2
n

n
= n
t
S +
1
2
n[S[
2
=
t
(nS) divJ S +
1
2
n[S[
2
=
t
J + div
_
J J
n
_
,
since, with the Hessian D
2
S,
1
2
n[S[
2
= n(D
2
S)S = div(nS S) div(nS)S
= div
_
J J
n
_
+ divJ S.
Proof of Theorem 6.1: (1) We can write the complex-valued wave function as =

n exp(iS/), where n = [[
2
and S is some phase function. Then
J = Im() = Im
_

n +
i

nS
_
= nS.
Thus, n and S satisfy the initial conditions (6.7). It remains to show that (n, J) solves
(6.8) since this implies (6.4)(6.5).
Inserting the expression =

n exp(iS/) into (6.2) gives, after division of exp(iS/),


i
2

t
n

n
t
S =

2
2
_

n +
2i

n S +
i

nS

2
[S[
2
_

nV.
137
The imaginary part of this equation equals

t
n = 2

n S nS = div(nS),
which is the rst equation in (6.8). The real part of the above equation reads

t
S =

2
2

n

1
2
[S[
2
+ V,
which is the second equation in (6.8).
(2) Dierentiation of =

n exp(iS/) gives, using (6.8),


i
t
+

2
2
= e
iS/
_
i

t
n
2

n
t
S +

2
2

n + i

n S
+
i
2

nS

n
2
[S[
2
_
= e
iS/
_

i
2
div(nS)

n
+ i

n S +
i
2

nS

nV
_
=

n e
iS/
V = V .

The system (6.4)(6.5) is the quantum analogue of the classical pressureless Euler
equations of gas dynamics. Notice that the above derivation requires an irrational initial
velocity J/n (since curl (J/n) = curl (S) = 0).
The quantum term can be interpreted either as a quantum self-potential term with
the so-called Bohm potential

n/

n or as a non-diagonal pressure tensor


P =

2
4
n() log n
since divP = (
2
/2)n(

n/

n).
6.2 Quantum hydrodynamics and the Schrodinger equation
The quantum hydrodynamic model of the previous section is derived for a single particle
and therefore, it does not contain a temperature term. In order to include temperature,
many-particle systems need to be studied. We assume that the considered electron ensem-
ble is represented by a mixed state (see Section 5.1). A mixed quantum state consists of
a sequence of single states with occupation probability
k
0, k N, for the k-th state
described by the Schrodinger equations
i
t

k
=

2
2

k
V
k
, x R
d
, t > 0, (6.9)

k
(x, 0) =
I,k
(x), x R
d
. (6.10)
138
The occupation probabilities clearly satisfy

k=1

k
= 1, i.e, the probability of nding
the electron ensemble in any of the quantum states is one.
Dene the single-particle particle and current densities of the k-th state as in the
previous subsection as
n
k
= [
k
[
2
, J
k
= Im(
k

k
), k N.
We claim that the total carrier density n and current density J of the mixed state, given
by
n =

k=1

k
n
k
, J =

k=1

k
J
k
, (6.11)
is a solution of the quantum hydrodynamic equations with temperature tensor:

t
n divJ = 0, (6.12)

t
J div
_
J J
n
+ n
_
+ nV +

2
2
n
_

n
_
= 0, x R
d
, t > 0, (6.13)
with initial conditions
n(, 0) =

k=1

k
[
I,k
[
2
, J(, 0) =

k=1

k
Im(
I,k

I,k
), in R
d
, (6.14)
where R
dd
is the temperature tensor dened below (see (6.15)).
Theorem 6.2 Let
k
be a solution of (6.9)(6.10) for all k N. Then (n, J), dened in
(6.11), solves (6.12)(6.14), where
=
c
+
os
(6.15)
and the current temperature
c
and osmotic temperature
os
are given by

c
=

k=1

k
n
k
n
(u
c,k
u
c
) (u
c,k
u
c
), (6.16)

os
=

k=1

k
n
k
n
(u
os,k
u
os
) (u
os,k
u
os
), (6.17)
and
u
c
=
J
n
, u
c,k
=
J
k
n
k
, u
os
=

2
ln n, u
os,k
=

2
ln n
k
(6.18)
are the current and osmotic velocities, respectively.
The notion osmotic comes from the fact that the quantum term can be written as

2
2
n
_

n
_
= divP with P =

2
4
n() ln n,
and P can be interpreted as a non-diagonal pressure tensor.
139
Proof: Theorem 6.1 shows that (n
k
, J
k
) solves the system of equations

t
n
k
divJ
k
= 0,

t
J
k
div
_
J
k
J
k
n
k
_
+n
k
V +

2
2
n
k

n
k

n
k
_
= 0, x R
d
n
k
(, 0) = [
I,k
[
2
, J
k
(, 0) = Im(
I,k

I,k
).
Multiplication of the above equations by
k
and summation over k yields

t
n divJ = 0,

t
J

k=1

k
div
_
J
k
J
k
n
k
_
+nV +

2
2

k=1

k
n
k

n
k

n
k
_
= 0. (6.19)
We rewrite the second and fourth term of the second equation. Using (6.16) and (6.18)
we obtain

k=1

k
div
_
J
k
J
k
n
k
_
(6.20)
=

k=1

k
div(n
k
u
c,k
u
c,k
)
=

k=1

k
div (n
k
(u
c,k
u
c
) (u
c,k
u
c
) + 2n
k
u
c,k
u
c
) div(nu
c
u
c
)
= div(n
c
) +

k=1
div
_
2
k
J
k
J
n
_
div
_
J J
n
_
= div(n
c
) + div
_
J J
n
_
. (6.21)
Furthermore, employing (6.17) and (6.18),

k=1

k
n
k

n
k

n
k
_
=
1
2

k=1

k
div
_
()n
k

n
k
n
k
n
k
_
=
1
2

k=1

k
div
_
()n
k
+
n n
n
2
2
n n
k
n
n
k
_
n
k
n
k

n
n
_

_
n
k
n
k

n
n
__
=
1
2
div
_
()n
n n
n
_
+ div(n
os
)
= n
_

n
_
+ div(n
os
). (6.22)
140
Putting (6.21) and (6.22) into (6.19) gives (6.13).
The temperature tensor cannot be expressed in terms of the particle and current den-
sities without further assumptions, and as in the derivation of the classical hydrodynamic
model (see Section 4.2), we need a closure condition to obtain a closed set of equations.
In the literature, the following conditions have been used:
Derive an evolution equation for the temperature tensor and close the corresponding
equation (we do not consider this approach here since we derive a temperature
equation in Section 6.3).
Assume that the temperature tensor is diagonal with equal entries on the diagonal,
= T Id ,
where T is a scalar temperature and Id R
dd
is the identity matrix. Usually,
the temperature T is a constant (isothermal case) but also a nonlinear temperature
T(n) = T
0
n
1
for some > 1 can be considered (isentropic case).
The second choice implies that the (classical) pressure n becomes
n = T
0
n

Id ,
with = 1 in the isothermal case and > 1 in the isentropic case.
6.3 Quantum hydrodynamics and the Wigner equation
The quantum hydrodynamic model of the previous section does not include any collisional
eects. In fact, the Schrodinger equation only models ballistic quantum transport. In
order to include scattering phenomena, we use the Wigner equation of Section 5.3 with
the Caldeira-Leggett collision operator (see (5.32)(5.33)):

t
w +
k
m


x
w +
q
m

[V ]w =
1

0
div
k
_
kw +
m

k
B
T
0

2

k
w
_
, t > 0, (6.23)
w(x, k, 0) = w
I
(x, k), x, k R
d
,
where we recall the denition of the pseudodierential operator [V ],
([V ]w)(x, k, t) =
i
(2)
d
_
R
d
_
R
d
m

_
V
_
x +

2
, t
_
V
_
x

2
, t
__
w(x, k
t
, t)e
i(kk

)
dk
t
d,
141
and m

is the eective electron mass. According to (5.31) we have assumed a parabolic


energy band E(k) =
2
[k[
2
/2m

, implying
v(k) =
1

k
E(k) =
k
m

.
First we scale the so-called Wigner-Fokker-Planck equation (6.23). For this we intro-
duce reference values for the time , the length L, and the potential U by requiring that
k
B
T
0
= m

_
L

_
2
and m

_
L

_
2
= qU.
This means that we consider a scaling in which thermal, kinetic, and electric energy are
of the same order. Then the scaling
t
s
=
t

, x s =
x
L
, k
s
=

m

L
k, V
s
= (x
s
, t
s
) =
V (x, t)
U
leads, after some elementary computations, to the following dimensionless equations (omit-
ting the index s):

t
w + k
x
w + [V ]w = div
k
(kw) +
k
w, t > 0, (6.24)
w(x, k, 0) = w
I
(x, k) x, k R
d
,
with the scaled operator
([V ]w)(x, k, t) =
i
(2)
d
_
R
d
_
R
d
1

_
V
_
x +

2
, t
_
V
_
x

2
, t
__
w(x, k
t
, t)e
i(kk

)
dk
t
d.
The parameter
=
/
m

(L/)
2
denotes the scaled Planck constant.
In order to derive macroscopic equations from (6.24) we apply the moment method as
in Section 4.2. We dene similarly as in the classical case the electron and current density
by
n(x, t) =
_
R
d
w(x, k, t) dk, J(x, t) =
_
R
d
w(x, k, t)k dk,
and we set
f(k)) =
_
R
d
w(x, k, t)f(k) dk
for any function f. Then n = 1) and J = k).
142
Lemma 6.3 The rst moment equations corresponding to (6.24) are as follows:

t
n div
x
J = 0, (6.25)

t
J div
x
k k) +n
x
V = J, (6.26)

1
2
[k[
2
) + div
x

1
2
k[k[
2
) +
x
V J = [k[
2
) + dn, (6.27)
where x R
d
, t > 0. The initial conditions are given by
n(x, 0) =
_
R
d
w
I
(x, k) dk,
J(x, 0) =
_
R
d
w
I
(x, k)k dk,

1
2
[k[
2
)(x, 0) =
1
2
_
R
d
w
I
(x, k)[k[
2
dk, x R
d
.
Proof: First we integrate (6.24) over R
d
with respect to k:

t
1) + div
x
k) +
_
R
d
[V ]wdk =
_
R
d
div
k
(kw +
k
w) dk = 0. (6.28)
For the evaluation of the integral on the left-hand side we recall the Fourier transform

(x, , t) =
1
(2)
d/2
_
R
d
(x, k, t)e
ik
dk
with inverse
(x, k, t) =
1
(2)
d/2
_
R
d

(x, , t)e
ik
d.
Integrating the last equation over R
d
with respect to k gives the formula

(x, 0, t) =
1
(2)
d
_
R
d
_
R
d

(x, , t)e
ik
d dk. (6.29)
With this expression we compute
_
R
d
[V ]wdk =
i
(2)
d/2
_
R
d
_
R
d
1

_
V (x +

2
, t) V (x

2
, t)
_
(6.30)
w(x, , t)e
ik
d dk
= (2)
d/2
i

_
V (x +

2
, t) V (x

2
, t)
_
=0
w(x, 0, t)
= 0. (6.31)
Therefore, (6.28) equals (6.25).
143
Now we multiply (6.24) with k
j
and integrate over R
d
with respect to k. We compute
rst, integrating by parts,
_
R
d
div
k
(kw +
k
w)k
j
dk =
_
R
d
(kw +
k
w)
k
k
j
dk
=
_
R
d
_
k
j
w +
w
k
j
_
dk = k
j
) = J
j
and, using (6.29),
_
R
d
[V ]wk
j
dk (6.32)
=
i
(2)
d/2
_
R
d
_
R
d
1

_
V (x +

2
, t) V (x

2
, t)
_
w(x, , t)(i)

j
e
ik
d dk
= (2)
d/2
1

j
_
V (x +

2
, t) V (x

2
, t)
_
=0
w(x, 0, t)
(2)
d/2
1

_
V (x +

2
, t) V (x

2
, t)
_
=0
w

j
(x, 0, t)
=
1
2
_
V
x
j
(x +

2
, t) +
V
x
j
(x

2
, t)
_
=0
_
R
d
w(x, k, t) dk
=
V
x
j
n(x, t). (6.33)
Thus

t
k
j
) + div
x
k
j
k) n
V
x
j
= J
j
.
This equals (6.26).
Finally, we multiply (6.24) with
1
2
[k[
2
and integrate over k R
d
. Again, we rst
compute the collision and electric-force terms. We obtain, integrating by parts,
1
2
_
R
d
div
k
(kw +
k
w)[k[
2
dw =
_
R
d
(kw +
k
w) k dk
=
_
R
d
(w[k[
2
+wdiv
k
k) dk = [k[
2
) + dn
144
and, with the notation V (x, , t) = [V (x +/2, t) V (x /2, t)]/ and using (6.29),
1
2
_
R
d
[V ]w[k[
2
dk =
i
2(2)
d/2
_
R
d
_
R
d
V (x, , t) w(x, , t)(

)e
ik
d dk
=
i
2(2)
d/2
_
R
d
_
R
d

(V w)(x, , t)e
ik
d dk
=
i
2
(2)
d/2

(V w)(x, 0, t)
=
i
2
(2)
d/2
_

V (x, 0, t) w(x, 0, t) + 2

V (x, 0, t)

w(x, 0, t)
+V (x, 0, t)

w(x, 0, t)
_
=
x
V k) =
x
V J,
since
V (x, 0, t) = 0,

V (x, 0, t) =
1
2
_

x
V (x +

2
, t) +
x
V (x

2
, t)
_
=0
=
x
V,
5

V (x, 0, t) =

4
_

x
V (x +

2
, t)
x
V (x

2
, t)
_
=0
= 0,

w(x, 0, t) =
1
(2)
d/2
_
R
d
w(x, k, t)

e
ik
dk

=0
=
i
(2)
d/2
k).
Thus, (6.27) is shown.
The system of moment equations (6.25)(6.27) has to be closed. As in the case of
classical kinetic theory, we achieve the closure by assuming that the Wigner function w
is close to a wave vector displaced equilibrium density,
w(x, k, t) = w
eq
(x, k u(x, t), t),
similar to (4.43), where w
eq
is assumed to be given by
w
eq
(x, k, t) = A(x, t)e
([k[
2
/2+V )/T(x,t)
_
1 +

2
8T(x, t)
2

x
V +

2
24T(x, t)
3
[
x
V [
2


2
24T(x, t)
3
d

j,=1
k
j
k

2
V
x
j
x

+ O(
4
)
_
, (6.34)
and the functions u(x, t) and T(x, t), later interpreted as the electron group velocity and
the electron temperature, have to be determined. Notice that the equilibrium density w
eq
reduces in the classical case 0 to the classical Maxwellian,
w
cl
eq
(x, k, t) = A(x, t)ne
[k[
2
/2T
,
145
since n = e
V/T
in classical thermal equilibrium. The rst moments with respect to w
eq
are
computed in the following lemma.
Lemma 6.4 The rst moments corresponding to (6.34) are
n = 1) = B(x, t)e
V/T
_
1 +

2
12T
2

x
V +

2
24T
3
[
x
V [
2
_
+ O(
4
), (6.35)
J = k) = nu, (6.36)
k k) = nu u + nT Id

2
12T
n()V + O(
4
), (6.37)
k[k[
2
) = nu[u[
2
+
_
dT

2
12T

x
V
_
nu + 2
_
T Id

2
12T
()V
_
nu +O(
4
),
(6.38)
where
B(x, t) = A(x, t)(2T)
d/2
,
Id is the identiy matrix of R
dd
and ()V the Hessian of V .
Proof: The symmetry of w
eq
with respect to k implies that all odd order moments of w
eq
vanish. Therefore,
1) =
_
R
d
w
eq
(x, k u(x, t), t) dk =
_
R
d
w
eq
(x, , t) d,
k) =
_
R
d
( + u)w
eq
(x, , t) d = nu,
k k) =
_
R
d
( + u) ( + u)w
eq
(x, , t) d = nu u +
_
R
d
w
eq
d, (6.39)
k[k[
2
) = nu[u[
2
+u
_
R
d
[[
2
w
eq
d + 2
__
R
d
w
eq
d
_
u. (6.40)
We derive now an O(
2
) approximation for n. Since
_
R
d
e
([[
2
/2+V )/T
d = T
d/2
e
V/T
_
R
d
e
[z[
2
/2
dz = (2T)
d/2
e
V/T
,
_
R
d

e
([[
2
/2+V )/T
d = T
d/2+1
e
V/T
_
R
d
z
j
z

e
[z[
2
/2
dz
= (2T)
d/2
Te
V/T
Id ,
we obtain
n = Be
V/T
_
1 +

2
8T
2

x
V +

2
24T
3
[
x
V [
2


2
24T
3
d

j,=1

2
V
x
j
x

j
_
+O(
4
)
= Be
V/T
_
1 +

2
12T
2

x
V +

2
24T
3
[
x
V [
2
_
+ O(
4
). (6.41)
146
The computation of the integrals on the right-hand sides of (6.39)(6.40) requires the
evaluation of
_
R
d

n
e
[[
2
/2T
d = T
d/2+2
_
R
d
z
j
z

z
m
z
n
e
[z[
2
/2
dz
= T
d/2+2
(j, , m, n),
where
(j, , m, n) = (2)
d/2
_

_
3 : j = = m = n
1 : j = ,= m = n or j = m ,= = n or j = n ,= = m
0 : else.
This yields
_
R
d

w
eq
d = Be
V/T
_
T
_
1 +

2
8T
2

x
V +

2
24T
2
[
x
V [
2
_
d
j


2
24T
3
d

m,n=1
T
2
(j, , m, n)

2
V
x
m
x
n
_
+ O(
4
).
A simple computation shows that
d

m,n=1
(j, , m, n)

2
V
x
m
x
n
= 2

2
V
x
j
x

+
x
V
j
.
Hence
_
R
d

w
eq
d = Be
V/T
_ _
T +

2
12T

x
V +

2
24T
2
[
x
V [
2
_
d
j


2
12T

2
V
x
j
x

_
+ O(
4
). (6.42)
We substitute the following expressions, derived from (6.41),
Be
V/T
_
T +

2
12T

x
V +

2
24T
2
[
x
V [
2
_
= nT + O(
4
),
Be
V/T
= n + O(
2
),
into (6.42) yielding
_
R
d

w
eq
d = nT
j


2
12T
n

2
V
x
j
x

+ O(
4
).
Putting this into (6.39) and (6.40) gives the remaining formulas (6.37) and (6.38).
147
We wish to replace the second derivatives of the electrostatic potential in (6.37) and
(6.38) by (derivatives of) the electron density. From the Taylor expansion log(x + ) =
log x + O() and the formula (6.35), written as
n = B(x, t)e
V/T
+ O(
2
),
we obtain
ln n = ln(B(x, t)e
V/T
) + O(
2
)
and therefore

2
ln n
x
j
x

=

2
x
j
x

_
ln B(x, t) +
V
T
_
+ O(
2
).
Our main assumption is that B(x, t) and T(x, t) are slowly varying with respect to x,
compared to spatial variations of V (x, t) such that we can write

2
ln n
x
j
x

=
1
T

2
V
x
j
x

+O(
2
). (6.43)
Lemma 6.5 Under the assumption (6.43) the second and third moments corresponding
to (6.34) are
k k) =
J J
n
+ nT Id

2
12
n() ln n + O(
4
),
k[k[
2
) =
J[J[
2
n
2

_
dT

2
12

x
ln n
_
J 2
_
T Id

2
12
() ln n
_
J + O(
4
).
Proof: The lemma immediately follows from (6.37)(6.38) and (6.43) using J = nu.
Now we can combine Lemmas 6.3 and 6.5 to achieve our main result.
Theorem 6.6 The rst moment equations corresponding to (6.24) with the closure con-
dition (6.34) and the assumption (6.43) are given by, up to order O(
4
),

t
n divJ = 0, (6.44)

t
J div
J J
n
(nT) +nV +

2
6
n
_

n
_
= J, (6.45)

t
e div
_
(e P)
J
n
_
+V J = 2
_
e
d
2
n
_
, (6.46)
where
e =
[J[
2
2n
+
d
2
nT

2
24
nln n,
P = nT Id +

2
12
n() ln n
are the energy density and the stress tensor, respectively.
148
Equations (6.44)(6.46) are solved for x R
d
, t > 0, supplemented with the initial
condtions
n(, 0) = n
I
, J(, 0) = J
I
, e(, 0) = e
I
in R
d
.
Proof: A simple computation shows that
div (n() ln n) = 2n
_

n
_
,
which proves (6.45). Equation (6.46) follows from (6.27) and Lemma 6.5.
Equations (6.44)(6.46) are called the quantum hydrodynamic equations. The above
energy density is the sum of the kinetic energy, [J[
2
/2n = n[u[
2
/2, thermal energy,
(d/2)nT, and quantum energy, (
2
/24)nln n. The stress tensor consists of a classi-
cal part, nT Id , and a quantum part, (
2
/12)n() ln n. It allows to write (6.45) in
the more compact form

t
J div
_
J J
n
+ P
_
+ nV = J.
We have used the formulation (6.45) since it allows to interpret the quantum term as a
force term with the self-interacting potential

n/

n, which is also called the Bohm


potential.
The terms on the right-hand sides of (6.45) and (6.46) are called the momentum and
energy relaxation terms, respectively. Neglecting all forces in the equations, we obtain

t
J = J,
t
e = 2
_
e
d
2
n
_
.
The solution to this system satises
J(t) 0, e(t)
d
2
n,
i.e., the current density vanishes and the energy relaxes to the thermal energy (with
constant temperature T = 1). We notice that for constant temperature, equations (6.44)
(6.45) give the quantum hydrodynamic model from Section 6.2 if = T.
Finally, we rewrite the quantum hydrodynamic equations in unscaled form:

t
n
1
q
divJ = 0,

t
J
1
q
div
_
J J
n
_

qk
B
m

(nT) +
q
2
m

nV +
q
2
6(m

)
2
n
_

n
_
=
J

t
e
1
q
div
_
(e P)
J
n
_
+ J V =
2

_
e
d
2
nT
0
_
,
149
where now
e = m

[J[
2
2n
+
d
2
nk
B
T

2
24m

ln n,
P = nk
B
T Id +

2
12m

n() ln n.
6.4 Derivation of the quantum drift-diusion equations
In this section we derive a quantum analogue of the classical drift-diusion equations from
a Wigner-Boltzmann model (i.e. the Wigner equation including a collision operator). Our
main assumption is that collisions are dominant, similar to the situation of Section 4.1.
We proceed as in [25] with the scaled Wigner equation

t
w + k
x
w + [V ]w = Q(w), t > 0, (6.47)
w(x, k, 0) = w
I
(x, k), x, k R
d
(6.48)
(see (6.24)), recalling the denition of the pseudodierential operator [V ],
([V ]w)(x, k, t) =
i
(2)
d
_
R
d
_
R
d
V (x, , t)w(x, k
t
, t)e
i(kk

)
dk
t
d,
where
V (x, , t) =
1

_
V (x +

2
, t) V (x

2
, t)
_
.
The collision operator Q(w) is assumed to be of relaxation type, i.e.
Q(w) = M[w] w, (6.49)
where M[w] is a quantum analogue of the Maxwellian exp(A(x, t) [k[
2
/2) for some
function A(x, t). The quantum Maxwellian M[w] is assumed to minimize the quantum
entropy, under the constraint of given particle density.
More precisely, similar to Section 4.2 we introduce the quantum entropy
H(w) =
_
R
d
_
R
d
w
_
Ln w 1 +
[k[
2
2
V (x)
_
dxdk,
where here and in the following we omit the time variable which only acts as a parameter
and where the quantum logarithm is dened by
Ln w = W(ln W
1
(w)).
In this denition, W is the Wigner transform
(W())(x, k) =
1
(2)
d
_
R
d

_
x +

2
, x

2

_
e
ik
d
150
and W
1
its inverse, the so-called Weyl quantization,
(W
1
(w))(x) =
_
R
d
_
R
d
w
_
x + y
2
, k
_
(y)e
ik(xy)/
dk dy,
where (x, x
t
) is a density matrix and w(x, k) a function. The logarithm ln is dened in
the sense of operator theory.
The quantum logarithm has the following properties.
Lemma 6.7 (1) The inverse of the quantum logarithm Ln is given by the so-called
quantum exponential Exp, dened by
Exp w = W(exp W
1
(w)) for suitable functions w(x, k).
(2) The functionals Ln and Exp are (Frechet) dierentiable and
d
dw
Ln w =
1
w
,
d
dw
Exp w = Exp w.
Proof: (1) Formally, it holds
Ln(Exp w) = W
_
exp W
1
_
W
_
ln W
1
(w)
_
__
= W
_
exp(ln W
1
(w))
_
= W(W
1
(w)) = w.
(2) Left to the reader.
Lemma 6.8 The solution of the constrained minimization problem
H(w

) = min
_
H(w) :
_
w(x, k) dk = n(x) for all x
_
,
for given electron density n(x), if it exists, is
w

= Exp(A(x) [k[
2
/2),
where the constant A(x) is xed by the constraint
_
R
d
w

(x, k) dk = n(x).
We assume that the above constraint xes A(x) uniquely.
Proof: We use Lagrange multipliers and dene
F(w, ) = H(w) +
_
R
d
(x)
__
R
d
w(x, k) dk n(x)
_
dx.
151
By Lemma 6.7 (2),
_
d
dw
H(w)
_
(u) =
_
R
d
_
R
d
_
Ln w +
[k[
2
2
V (x)
_
u(x, k) dxdk.
Furthermore,
_
d
dw
_
R
d
wdk
_
(g) =
_
R
d
g(k) dk.
Therefore, we obtain the necessary condition for an extremum
0 =
_
d
dw
F(w

)
_
(u) =
_
R
d
_
R
d
_
Ln w

+
[k[
2
2
V (x) +

(x)
_
u(x, k) dxdk
for all functions u(x, k). This implies that
Ln w

+
[k[
2
2
V (x) +

(x) = 0
or
w

= Exp(A(x) [k[
2
/2) for A(x) = V (x)

(x),
which shows the lemma.
Denition 6.9 The quantum Maxwellian is dened by
M[w] = Exp(A(x) [k[
2
/2)
and A(x) is determined through
_
R
d
M[w] dk =
_
R
d
wdk.
With this denition, the relaxation-time Wigner model (6.47)(6.49) is well dened.
In order to derive the quantum drift-diusion model, we consider the case of dominant
collisions and large time, i.e., we rescale the equation (6.47) by replacing
t by t/ and Q(w) by Q(w)/.
This yields

t
w

+ k
x
w

+ [V ]w

= Q(w

) =
1

(M[w

] w

), t > 0, (6.50)
w

(x, k, 0) = w
I
(x, k), x, k R
d
. (6.51)
Now we perform the limit 0.
152
Theorem 6.10 Let w

be a solution of (6.50)(6.51). Then, formally, w

w
0
as 0,
where w
0
= Exp(A [k[
2
/2) and A(x, t) is a solution of

t
n divJ = 0, J = divP nV, t > 0 (6.52)
n(x, 0) = n
I
(x), x R
d
, (6.53)
where
n =
_
R
d
Exp
_
A
[k[
2
2
_
dk, (6.54)
P =
_
R
d
k k Exp
_
A
[k[
2
2
_
dk, (6.55)
and the initial datum is given by
n
I
=
_
R
d
w
I
dk.
Proof: Performing the (formal) limit 0 in (6.50) directly, we obtain w
0
= M[w
0
]. In
order to derive higher-order corrections, we apply the so-called Chapman-Enskog method
rst used in classical kinetic theory [20, 29]. For this, we dene the rst-order correction
w
(1)

=
1

(w

M[w

]) or w

= M[w

] + w
(1)

.
Denote by w
(1)
the limit of w
(1)

. Since, as 0,
w
(1)

= Q(w

) =
t
w

(k
x
w

+[V ]w

) (k
x
w
0
+ [V ]w
0
),
we infer
w
(1)
= (k
x
w
0
+ [V ]w
0
). (6.56)
We claim that in the (formal) limit 0,

t
_
R
d
w
0
dk + div
x
_
R
d
w
(1)
k dk +
_
R
d
[V ]w
(1)
dk = 0. (6.57)
To see this we integrate the rescaled Wigner equation (6.50) with respect to k R
d
,

t
_
R
d
w

dk + div
x
_
R
d
w

k dk +
_
R
d
[V ]w

dk =
_
R
d
Q(w

) dk, (6.58)
and manipulate this equation term by term. By denition, M[w

] and w

have the same


(zeroth-order) moments, yielding
_
R
d
Q(w

) dk =
1

_
R
d
(M[w

] w

) dk = 0.
153
The function k ([V ]M[w

])(x, k) is odd since, transforming , k


t
k
t
,
([V ]M[w

])(x, k) =
i
(2)
d/2
_
R
2d
V (x, )Exp
_
A
[k
t
[
2
2
_
e
i(kk

)
d dk
t
=
i
(2)
d/2
_
R
2d
V (x, )Exp
_
A
[k
t
[
2
2
_
e
i(kk

)
d dk
t
= ([V ]M[w

])(x, k),
using V (x, ) = V (x, ). This implies that the integral over [V ]M[w

] vanishes and
therefore,
_
R
d
[V ]w

dk =
_
R
d
[V ](M[w

] + w
(1)

) dk =
_
R
d
[V ]w
(1)

dk.
Since also k M[w

](x, k)k is odd,


_
R
d
w

k dk =
_
R
d
w
(1)

k dk.
Thus (6.57) follows from (6.58) after division by .
Setting n =
_
w
0
dk and using the expression (6.56) in (6.57) we obtain

t
n div
x
_
R
d
(k
x
w
0
+[V ]w
0
)k dk +
_
R
d
[V ]w
(1)
dk = 0.
In Section 6.3 we have derived the following formulas (see (6.31) and (6.33)):
_
R
d
[V ]wdk = 0 for all functions w,
_
R
d
[V ]w
0
k dk = n
x
V for all functions w
0
such that n =
_
w
0
dk.
This gives

t
n div
x
_
R
d
div
x
(k k)w
0
dk + div
x
(nV
x
) = 0,
and hence the conclusion.
The initial-value problem (6.52)(6.55) can be considered as a generalized quantum
drift-diusion equation. The drift current is given by nV and P is a quantum pressure
tensor. In the classical case, P = n
ij
yielding the drift-diusion equations of Section 4.1.
In general, P is a nonlocal operator involving A(x) which relates to n through (6.54)
(6.55). The usual quantum drift-diusion model is obtained from (6.52)(6.55) as a O(
4
)
approximation.
154
Theorem 6.11 Let (n, J) be a solution of the generalized quantum drift-diusion equa-
tions (6.52)(6.55). Then, formally, J = J
0
+ O(
4
) and (n, J
0
) solves

t
n divJ
0
= 0, J
0
= n + nV +

2
6
n
_

n
_
, t > 0, (6.59)
n(x, 0) = n
I
(x), x R
d
. (6.60)
The system (6.59)(6.60) is called the quantum drift-diusion model. For the proof of
Theorem 6.11 we need the following results. It holds
_
R
d
e
[k[
2
/2
dk = (2)
d/2
, (6.61)
_
R
d
k
i
k
j
e
[k[
2
/2
dk = (2)
d/2

ij
, (6.62)
_
R
d
k
i
k
j
k

k
m
e
[k[
2
/2
dk = (2)
d/2
(
ij

m
+
i

jm
+
im

j
) . (6.63)
Lemma 6.12 The following (formal) expansion holds:
Exp
_
A
[k[
2
2
_
= exp
_
A
[k[
2
2
__
1 +

2
8
_
A +
1
3
[A[
2

1
3
k

D
2
Ak
__
+ O(
4
),
where D
2
A denotes the Hessian of A.
The technical proof uses heavily pseudo-dierential calculus and is therefore omitted.
We refer to [25, Lemma 5.6] for details of the proof.
Proof of Theorem 6.11: First we develop the electron density in powers of
2
. By (6.54),
Lemma 6.12, and (6.61)(6.62),
n = e
A
_
1 +

2
8
_
A +
1
3
[A[
2
__ _
R
d
e
[k[
2
/2
dk

2
24
e
A

i,j

2
ij
A
_
R
d
k
i
k
j
e
[k[
2
/2
dk +O(
4
)
= (2)
d/2
e
A
_
1 +

2
8
_
A +
1
3
[A[
2
_


2
24
A
_
+O(
4
)
= (2)
d/2
e
A
_
1 +

2
12
_
A +
1
2
[A[
2
__
+ O(
4
), (6.64)
where
i
abbreviates the partial derivative /x
i
. In particular,
n = (2)
d/2
e
A
+O(
2
). (6.65)
155
Next we develop the quantum pressure in powers of
2
. Using (6.55), Lemma 6.12, and
(6.62)(6.63),
P
ij
= e
A
_
1 +

2
8
_
A +
1
3
[A[
2
__ _
R
d
k
i
k
j
e
[k[
2
/2
dk

2
24
e
A

,m

2
m
A
_
R
d
k
i
k
j
k

k
m
e
[k[
2
/2
dk + O(
4
)
= (2)
d/2
e
A
_
1 +

2
8
_
A +
1
3
[A[
2
__

ij


2
24
(2)
d/2
e
A
_
A
ij
+ 2
2
ij
A
_
+ O(
4
)
= (2)
d/2
e
A
_
1 +

2
12
_
A +
1
2
[A[
2
__

ij


2
12
(2)
d/2
e
A

2
ij
A + O(
4
).
We employ the developments (6.64) and (6.65) to obtain
P
ij
= n
ij


2
n
12

2
ij
A + O(
4
).
Thus, using n = nA + +O(
2
), which follows immediately from (6.65),
(divP)
j
=

i
P
ij
=
j
n

2
12

i
_

i
n
2
ij
A + n
3
iij
A
_
+O(
4
)
=
j
n

2
12

i
n
j
_
1
2
(
i
A)
2
+
ii
A
_
+ O(
4
).
Hence,
divP = n

2
12
n
_
A +
1
2
[A[
2
_
+ O(
4
).
Finally, we relate A and n using (6.65) or
A =
n
n
+O(
2
),
from which we conclude
A +
1
2
[A[
2
=
n
n

[n[
2
n
2
+
1
2

n
n

2
+ O(
2
) =
n
n

[n[
2
2n
2
+ O(
2
)
= 2

n
+ O(
2
).
Therefore,
J = divP nV = n nV

2
2
n
_

n
_
+ O(
4
),
and the conclusion follows.
156
References
[1] R. Adler, A. Smith, and R. Longini. Introduction to Semiconductor Physics. John
Wiley & Sons, New York, 1964.
[2] A. Arnold, P. Degond, P. Markowich and H. Steinr uck. The Wigner-Poisson equation
in a crystal. Appl. Math. Lett. 2 (1989), 187-191.
[3] N. Ashcroft and N. Mermin. Solid State Physics. Sanners College, Philadelphia, 1976.
[4] H. Babovsky. Die Boltzmann-Gleichung. Teubner, Stuttgart, 1998.
[5] G. Baccarani and M. Wordeman. An investigation on steady-state velocity overshoot
in silicon. Solid-State Electr. 29 (1982), 970-977.
[6] C. Bardos, F. Golse, and C. Levermore. Fluid dynamic limits of kinetic equations I:
formal derivations. J. Stat. Phys. 63 (1991), 323-344.
[7] C. Bardos, F. Golse, and C. Levermore. Fluid dynamic limits of kinetic equations II:
convergence proofs for the Boltzmann equation. Comm. Pure Appl. Math. 46 (1993),
667-753.
[8] N. Ben Abdallah and P. Degond. On a hierarchy of macroscopic models for semicon-
ductors. J. Math. Phys. 37 (1996), 3308-3333.
[9] N. Ben Abdallah, P. Degond, and S. Genieys. An energy-transport model for semi-
conductors derived from the Boltzmann equation. J. Stat. Phys. 84 (1996), 205-231.
[10] K. Bltekjr. Transport equations for electrons in two-valley semiconductors. IEEE
Trans. Electr. Devices 17 (1970), 38-47.
[11] K. Brennan. The Physics of Semiconductors. Cambridge University Press, Cam-
bridge, 1999.
[12] N. Bogoliubov. Problems of a dynamical theory in statistical physics. In: J. de Boer,
U. Uhlenbeck (eds.). Studies in Statistical Mechanics, Vol. I. North-Holland, Ams-
terdam (1962), 5.
[13] L. Boltzmann. Weitere Studien uber das Warmegleichgewicht unter Gasmolek ule.
Sitzungsberichte Akad. Wiss. Wien 66 (1872), 275-370. Translation: Further studies
on the thermal equilibrium of gas molecules. In: S. Brush (ed.). Kinetic Theory, Vol.
2. Pergamon Press, Oxford (1966), 88-174.
157
[14] M. Born and H. Green. A General Kinetic Theory of Fluids. Cambridge University
Press, Cambridge, 1949.
[15] F. Buer and A. Schenk. Halbleitertransporttheorie und Monte-Carlo-
Bauelementsimulation. Lecture Notes, ETH Z urich, Switzerland, 2002.
[16] A. Caldeira and A. Leggett. Path integral approach to quantum Brownian motion.
Physica A 121A (1983), 587-616.
[17] A. Caldeira, L. Erdos, F. Frommlet, and P. Markowich. Fokker-Planck equations as
scaling limits of reversible quantum systems. J. Stat. Phys. 100 (2000), 543-601.
[18] P. Carruthers and F. Zachariasen. Quantum collision theory with phase-space distri-
butions. Rev. Modern. Phys. 55 (1983), 245-285.
[19] C. Cercignani. The Boltzmann Equation and Its Applications. Springer, Berlin, 1988.
[20] S. Chapman. The kinetic theory of simple and composite gases: viscosity, thermal
conduction and diusion. Proc. Roy. Soc. London A 93 (1916/17), 1-20.
[21] D. Chen, E. Kan, U. Ravaioli, C. Shu, and R. Dutton. An improved energy trans-
port model including nonparabolicity and non-Maxwellian distribution eects. IEEE
Electr. Dev. Lett. 13 (1992), 26-28.
[22] R. Dautray and J.L. Lions. Mathematical Analysis and Numerical Methods for Sci-
ence and Technology. Springer, Berlin, 1992.
[23] P. Degond. Mathematical modelling of microelectronics semiconductor devices. In:
L. Hsiao (ed.). Some Current Topics on Nonlinear Conservation Laws. Amer. Math.
Soc., Providence, USA (2000), 77-100.
[24] P. Degond, A. J ungel, and P. Pietra. Numerical discretization of energy-transport
models for semiconductors with nonparabolic band structure. SIAM J. Sci. Comp.
22 (2000), 986-1007.
[25] P. Degond, F. Mehats, and C. Ringhofer. Quantum energy-transport and drift-
diusion models. J. Stat. Phys. 118 (2005), 625-665.
[26] R. DiPerna, P.L. Lions. On the Cauchy problem for Boltzmann equations: global
existence and weak stability. Ann. Math. 130 (1989), 321-366.
[27] R. DiPerna, P.L. Lions. Global solutions of Boltzmanns equation and the entropy
inequality. Arch. Ration. Mech. Anal. 114 (1991), 47-55.
158
[28] P. Dmitruk, A. Saul, and L. Reyna. High electric-eld approximation in semiconduc-
tor devices. Appl. Math. Letters 5 (1992), 99-102.
[29] D. Enskog. Kinetische Theorie der Vorgange in maig verd unnten Gasen. 1. Allge-
meiner Teil. Almquist & Wiksell, Uppsala, 1917.
[30] H. Federer. Geometric Measure Theory. Springer, New York, 1969.
[31] D. Ferry and H. Grubin. Modelling of quantum transport in semiconductor devices.
Solid State Phys. 49 (1995), 283-448.
[32] H. Foll. Semiconductors. Lecture Notes, Universitat Kiel, Germany, 2004.
(www.tf.uni-kiel.de/matwis/amat/semi en/index.html).
[33] V. Gantmakher and Y. Levinson. Carrier Scattering in Metals and Semiconductors.
North Holland, New York, 1987.
[34] A. Gnudi, D. Ventura, G. Baccarani, and F. Odeh. Two-dimensional MOSFET sim-
ulations by means of a multidimensional spherical harmonic expansion of the Boltz-
mann transport equation. Solid State Electr. 36 (1993), 575-581.
[35] N. Goldsman, L. Henrickson, and J. Frey. A physics based analytical numerical so-
lution to the Boltzmann transport equation for use in device simulation. Solid State
Electr. 34 (1991), 389-396.
[36] F. Golse. The Boltzmann Equation and Its Hydrodynamic Limits. Lecture Notes for
the Ecole thematique Les mod`eles cinetiques non lineaire et leurs limites uides,
Universite de Franche-Comte, Besan on, 1997.
[37] H. Grahn. Introduction to Semiconductor Physics. World Scientic, Singpore, 1999.
[38] R. Hudson. When is the Wigner quasiprobability nonnegative?, Reports Math. Phys.
6 (1974), 249-252.
[39] J. Kirkwood. Statistical mechanical theory of transport processes I. General theory.
J. Chem. Phys. 14 (1946), 180-201. Errata: 14, 347.
[40] C. Kittel. Introduction to Solid State Physcis. 7th edition, John Wiley & Sons, New
York, 1996.
[41] C. Lent and K. Kirkner. The quantum transmitting method. J. appl. Phys. 67 (1990),
6353-6359.
159
[42] C. Levermore. Moment closure hierarchies for kinetic theories. J. Stat. Phys. 83
(1996), 1021-1065.
[43] C. Levermore. Entropy-based moment closures for kinetic equations. Transp. Theory
Stat. Phys. 26 (1997), 591-606.
[44] P. L. Lions. Global solutions of kinetic models and related problems. In: C. Cercig-
nani and M. Pulvirenti (eds.). Nonequilibrium Problems in Many-Particle Systems.
Lecture Notes in Math., Vol. 1551. Springer (1992), 58-86.
[45] M. Lundstrom. Fundamentals of Carrier Transport. 2nd edition, Cambridge Univer-
sity Press, Cambridge, 2000.
[46] J. Mal y, D. Swanson, and W. Ziemer. The coarea formula for Sobolev mappings.
Trans. Amer. Math. Soc. 355 (2003), 477-492.
[47] P. Markowich. On the equivalence of the Schrodinger and the quantum Liouville
equations. Math. Meth. Appl. Sci. 11 (1989), 459-469.
[48] P. Markowich and C. Ringhofer. An analysis of the quantum Liouville equation. Z.
Angew. Math. Mech. 69 (1989), 121-127.
[49] P. Markowich, C. Ringhofer, and C. Schmeiser. Semiconductor Equations. Springer,
Vienna, 1990.
[50] D. Neamen. Semiconductor Physics and Devices. 3rd edition, McGraw Hill, 2002.
[51] F. Poupaud. Etude mathematique et simulations numeriques de quelques equations
de Boltzmann. PhD thesis, Universite de Nice, France, 1988.
[52] F. Poupaud. On a system of nonlinear Boltzmann equations of semiconductor physics.
SIAM J. Appl. Math. 50 (1990), 1593-1606.
[53] F. Poupaud. Diusion approximation of the linear semiconductor Boltzmann equa-
tion: analysis of boundary layers. Asympt. Anal. 4 (1991), 293-317.
[54] F. Poupaud. Mathematical theory of kinetic equations for transport modelling in
semiconductors. In: B. Perthame (ed.). Advances in Kinetic Theory and Computing:
Selected Papers. World Scientic, Singapore (1994), 141-168.
[55] F. Poupaud and C. Ringhofer. Quantum hydrodynamic models in semiconductor
crystals. Appl. Math. Lett. 8 (1995), 55-59.
160
[56] M. Reed and B. Simon. Methods of Modern Mathematical Physics IV: Analysis of
Operators. Academic Press, New York, 1978.
[57] L. Reggiani (ed.). Hot Electron Transport in Semiconductors. Springer, Berlin, 1985.
[58] B. Ridley. Quantum Processes in Semiconductors. Clarendon Press, Oxford, 1982.
[59] C. Ringhofer. Computational methods for semiclassical and quantum transport in
semiconductor devices. Acta Numerica (1997), 485-521.
[60] C. Schmeiser and A. Zwirchmayr. Elastic and drift-diusion limits of electron-phonon
interaction in semiconductor. Math. Models Meth. Appl. Sci. 8 (1998), 37-53.
[61] K. Seeger. Semiconductor Physics. An Introduction. 6th edition, Springer, Berlin,
1977.
[62] M. Taylor. Pseudodierential Operators. Princeton University Press, Princeton, 1981.
[63] C. Villani. A review of mathematical topics in collisional kinetic theory. In: S. Fried-
lander and D. Serre. Handbook of Mathematical Fluid Dynamics, Vol. 1. Elsevier,
Amsterdam (2002), 71-305.
[64] W. Van Roosbroeck. Theory of ow of electron and holes in germanium and other
semiconductors. Bell Syst. Techn. J. 29 (1950), 560-607.
[65] W. Wenckebach. Essentials of Semiconductor Physics. John Wiley & Sons, Chich-
ester, 1999.
[66] E. Wigner. On the quantum correction for the thermodynamic equilibrium. Phy. Rev.
40 (1932), 749-759.
[67] J. Yvon. La theorie statistique des uides. Hermann, Paris, 1935.
[68] W. Zawadzki. Mechanics of electron scattering in semiconductors. in: T. Moss (ed.).
Handbook of Semiconductors. Vol. 1. North-Holland, New York (1982), Chapter 12.
161

You might also like