0% found this document useful (0 votes)
61 views49 pages

Carrier-Transport Equations: Karl W. Böer and Udo W. Pohl

This document discusses carrier transport equations in semiconductors. It describes how electrons and holes in semiconductors interact with the lattice as quasi-particles known as polarons. Their motion is subjected to scattering, which reduces their effective mean free path. Both elastic and inelastic scattering events can occur as carriers gain or lose energy from electric fields. The total current in a semiconductor is composed of drift and diffusion currents of electrons and holes. Key concepts like conductivity, mobility, the Boltzmann equation, and scattering mechanisms are introduced to describe carrier transport.

Uploaded by

raul
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
61 views49 pages

Carrier-Transport Equations: Karl W. Böer and Udo W. Pohl

This document discusses carrier transport equations in semiconductors. It describes how electrons and holes in semiconductors interact with the lattice as quasi-particles known as polarons. Their motion is subjected to scattering, which reduces their effective mean free path. Both elastic and inelastic scattering events can occur as carriers gain or lose energy from electric fields. The total current in a semiconductor is composed of drift and diffusion currents of electrons and holes. Key concepts like conductivity, mobility, the Boltzmann equation, and scattering mechanisms are introduced to describe carrier transport.

Uploaded by

raul
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 49

Carrier-Transport Equations

Karl W. Böer and Udo W. Pohl

Contents
1 Carriers in Semiconductors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2
1.1 Bloch Electrons and Holes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
1.2 The Polaron . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
2 Conductivity and Mobility of Carriers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12
2.1 Electronic Conductivity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
2.2 Electron Mobility . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16
3 Currents and Electric Fields . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18
3.1 Drift Current in an Electric Field . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18
3.2 Diffusion Currents and Total Currents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21
3.3 Electrochemical Fields and Quasi-Fermi Levels . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25
3.4 Carrier Distributions in External and Built-In Fields . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28
4 The Boltzmann Equation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32
4.1 The Boltzmann Equation for Electrons . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33
4.2 The Boltzmann Equation for Phonons . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35
4.3 The Relaxation-Time Approximation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36
4.4 Carrier Scattering and Energy Relaxation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 38
4.5 The Mobility Effective Mass . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 41
4.6 Momentum and Energy Relaxation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43
4.7 Phonon and Electron Drag . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 46
5 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 46
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 48

K. W. Böer
Naples, FL, USA
e-mail: [email protected]
U. W. Pohl (*)
Institut für Festkörperphysik, EW5-1, Technische Universität Berlin, Berlin, Germany
e-mail: [email protected]

# Springer International Publishing AG 2017 1


K. W. Böer, U. W. Pohl, Semiconductor Physics,
https://doi.org/978-3-319-6/978-3-319-06540-3_22-2
2 K. W. Böer and U. W. Pohl

Abstract
The current in semiconductors is carried by electrons and holes. Their lattice
polarization modifies the effective mass, expressed as a change to polarons. While
for large polarons the effect is small, semiconductors with narrow bands and large
lattice polarization show a significant effect described by small polarons. The total
current is composed of a drift and a diffusion current of electrons and holes. The drift
current is determined by the electric field, and the energy obtained by carrier
acceleration is given to the lattice by inelastic scattering, which opposes the energy
gain, causing a constant carrier-drift velocity and Joule’s heating. The diffusion
current is proportional to the carrier gradient up to a limit given by the thermal
velocity. Proportionality factor of both drift and diffusion currents is the carrier
mobility, which is proportional to a relaxation time and inverse to the mobility
effective mass. Currents are proportional to negative potential gradients, with the
conductivity as the proportionality factor. In spatially inhomogeneous semiconduc-
tors, both an external field, impressed by an applied bias, and a built-in field, due to
space-charge regions, exist. Only the external field causes carrier heating by shifting
and deforming the carrier distribution from a Boltzmann distribution to a distorted
distribution with more carriers at higher energies.
The Boltzmann equation permits a detailed analysis of the carrier transport and the
carrier distribution, providing well-defined values for transport parameters such as
relaxation times. The Boltzmann equation can be integrated in closed form only for a
few special cases, but approximations for small applied fields provide the basis for
investigating scattering processes; these can be divided into essentially elastic pro-
cesses with mainly momentum exchange and, for carriers with sufficient accumu-
lated energy, into inelastic scattering with energy relaxation.

Keywords
Boltzmann transport equation  Built-in electric field  Carrier heating  Collision
integral  Conductivity  Diffusion current  Drift current  Drift velocity 
Effective mass  Einstein relation  Energy relaxation  Fröhlich coupling 
Inelastic and elastic scattering  Joule’s heating  Mean free path  Mobility
effective mass  Momentum relaxation  Polaron  Polaron mass  Polaron self-
energy  Quasi-Fermi level  Relaxation time  Relaxation-time approximation

1 Carriers in Semiconductors

Electrons and holes in the conduction and valence bands are quasi-free to move in space
and energy and to accept energy from an external field. Carrier transport in semi-
conductors with nonideal periodicity1 is subjected to scattering with a mean free path

1
A crystal with nonideal lattice periodicity is a solid which contains lattice defects (e.g., impurities)
and oscillatory motions (i.e., phonons).
Carrier-Transport Equations 3

λ between scattering events. Such scattering reduces the effective volume to a value of
λ3, in which there is coherence of the electron wave. Only within such limited volume
does any one electron experience lattice periodicity and is nonlocalized. Scattering
introduces a loss, i.e., a damping mechanism counteracting the energy gain from an
electric field. All scattering events change the carrier momentum, i.e., its direction of
motion. However, only some of them, the inelastic scattering events, significantly
change the energy of the carrier. Usually, several elastic scattering events are followed
by one inelastic event after the carrier has gained sufficient energy from the field to
permit inelastic scattering, usually by generating optical phonons. A large variety of
scattering events can be distinguished; they are listed in Sect. 4.4 and discussed in
chapter ▶ “Carrier Scattering at Low Electric Fields”. The sum of all of these determines
the carrier motion, its mobility, which will be defined later.
First a quasiclassical picture will be used in this chapter to describe the basic
elements of the carrier motion in a semiconductor. The quantum-mechanical part is
incorporated by using an effective mass rather than the rest mass, i.e., by dealing with
Bloch electrons or Bloch holes as quasiparticles. Before we discuss carrier transport, we
need to specify in a more refined model what we mean by carriers in semiconductors.
The excitation of electrons from the valence band (e.g., by absorption of photons)
creates a certain concentration n of electrons in the conduction bands. These
electrons interact with lattice imperfections, such as phonons, impurities, or other
deviations from an ideal periodicity. This interaction is termed a scattering event.
The scattering tends to bring the electrons into thermal equilibrium with the lattice
and, in doing so, to the lowest valley of the lowest conduction band.

1.1 Bloch Electrons and Holes

Near the bottom of the conduction band, the electron is described as a Bloch
electron,2 i.e., as an electron with an effective mass given by the curvature of
E(k), as defined in Sect. 2.2 of chapter ▶ “The Origin of Band Structure”, Eq. 20.
In an analogous description, the hole is described as a Bloch-type quasiparticle,
residing near the top of the uppermost valence bands, with an effective mass given
by their curvatures – see Sect. 1.1.1 of chapter ▶ “Bands and Bandgaps in Solids”.
The band picture, however, which is the basis for this discussion, results from a series
of approximations listed in chapter ▶ “Quantum Mechanics of Electrons in Crystals”.
The most severe one is the adiabatic approximation, which limits the electron-phonon
interaction. Within the band model, the Bloch electron interacts with a static potential of
the nuclei, while only the electrons surrounding each nucleus are polarized dynamically.
With sufficient coupling between the moving electron and the lattice (i.e., a sufficiently
large coupling constant ac – see Sect. 3 1.2.1, this is no longer justified. For these cases,

2
We are adopting here the picture of a localized electron. Such localization can be justified in each
scattering event. In this model, we use a gas-kinetic analogy with scattering cross-sections, e.g., for
electron-phonon interaction. This is equivalent to a description of the interaction of delocalized
electrons and phonons when calculating scattering rates.
4 K. W. Böer and U. W. Pohl

therefore, the Bloch electron picture needs to be augmented. The corresponding quasi-
particle derived from a higher approximation is the polaron.

1.2 The Polaron

The previous discussion of free carriers with an effective mass was based on an ideal
periodic lattice. There are many reasons why a real crystal lattice shows deviations
from this periodicity. The perturbations, which cause major changes in the carrier
trajectories, can be described as local scattering centers discussed in chapter ▶ “Car-
rier Scattering at Low Electric Fields”.
Another interaction that steadily accompanies the carrier throughout its entire
motion,3 however, is better incorporated in the effective-mass picture. It also has an
influence on the scattering effectiveness and thus on the mobility, which will be
discussed in this section. This interaction involves lattice polarization and causes
deviations from the Bloch electron picture discussed above. Deviations are signifi-
cant in lattices with a large coupling of carriers to the lattice, such as in ionic crystals
(typically alkali halides), crystals with a strong ionic character and large bandgap
energy (e.g., transition-metal oxides), or organic crystals. Also many typical semi-
conductors contain at least a fraction of ionic bonding, especially compound semi-
conductors with higher iconicity (e.g., nitrides or II-VI compounds). Moreover, a
strong polarization of the lattice occurs also in the neighborhood of certain impurities
or in highly disordered semiconductors. In all these cases the interaction of electrons
with lattice polarization becomes important. Such interaction can be static (for an
electron at rest) or dynamic (accompanying a moving electron).
The interaction involves the part of the Hamiltonian not considered in the band
theory when describing the Bloch electron. This adiabatic approximation neglects
the interaction of electrons with the induced motion of lattice atoms. In order to
include the motion due to the lattice polarization by the electron, several approaches
can be taken, all of which relate to the coupling of the electron with the lattice. Such
coupling can be expressed as (Hayes and Stoneham 1984):

• Fröhlich coupling, i.e., via interaction with longitudinal optical phonons


• Deformation-potential coupling, i.e., via the electric field produced by the strain
field in acoustic oscillations
• Piezoelectric coupling, i.e., via the electric field produced by acoustic phonons in
piezoelectric semiconductors

The first coupling effect is dominant. It can be described as absorption or


emission of virtual LO phonons by the electron, which thereby lowers its eigenstate
– see, e.g., Comas and Mora-Ramos (1989).

3
This interaction is also commonly described as an electron-phonon interaction, however, of a
different kind than that responsible for scattering. As a lattice deformation relates to phonons, the
interaction of electrons with the lattice causing a specific deformation can formally be described by
continuously absorbing and emitting phonons (see the following sections).
Carrier-Transport Equations 5

a b
E

Estrain

e ε

Epol

Eion

Fig. 1 (a) Schematic of the polarization cloud in the vicinity of an electron (e) in a ionic lattice.
Gray circles indicate regular lattice sites. (b) Polaron energy Epol given by the sum of elastic strain
energy Estrain and electrical polarization energy Eion

Another approach to describe the interaction is a static description in which the


electron induces a shift of the surrounding ions due to its own Coulomb field
(Fig. 1a). The eigenstates of the electron in this Coulomb funnel are hydrogen-
like, similar to that of a hydrogen-like donor (Sect. 1 of chapter ▶ “Shallow-Level
Centers”); the ground state describes its self-energy. This lowers the energy of a
Bloch electron accordingly.
The energy lowering can be estimated from a simple consideration. The strain of
the lattice environment illustrated in Fig. 1a requires an elastic work

Estrain ¼ ðC1 =2Þ e2 ΔV; (1)

where e describes the relative displacements of the ions within the deformation
volume ΔV and the constant C1 is the energy density for a strain e = 1. This energy
cost is overbalanced by an energy gain Eion due to the polarization:

Eion ¼ C2 e ΔV: (2)

The constant C2 is an energy density, which describes the change of the charge
balance at a strain e = 1. The sum of Eqs. 1 and 2 yields the net energy gain Epol by
the formation of the polaron state as shown in Fig. 1b:

Epol =ΔV ¼ ðC1 =2Þ e2  C2 e; (3)

with a minimum C22 =2C1 at the strain e ¼ C2 =C1 .


6 K. W. Böer and U. W. Pohl

For a sufficiently large coupling to the lattice, the self-energy is large enough so
that the electron is actually trapped in its own potential well: it becomes a self-
trapped electron. This was suggested by Landau (1933) and Frenkel (1936) for bulk
crystals and later reported also in semiconductor confined systems such as quantum
wires (e.g., Muljarov and Tikhodeev 1997). Pekar (1954) pointed out, however, that
when the electron moves to a neighboring lattice position, the same trap level
appears. Thus, one can describe this as a continuous virtual level or a band below
the conduction band in which the electron can move. This picture applies for lattices
with smaller coupling constant.
The state of the electron with its surrounding polarization cloud can then be
described as a quasiparticle, the polaron. The distinguishing parameter between the
self-trapped and the mobile polaron is the strength of the electron-lattice interaction,
which also can be related to the polaron size. The small polaron is tightly bound and
self-trapped: it moves via hopping between neighboring ions. The large polaron
moves much like an electron described by the Boltzmann equation with scattering
events, but with a larger effective mass caused by the polarization cloud carried
along by the polaron (for a review, see Christov 1982). One can also describe the
polaron as an electron surrounded by a cloud of virtual phonons, which represents its
surrounding polarization. One refers to these polarons as phonon-dressed electrons,
since attached to them is the “fabric” of the surrounding lattice. They can be
described by a Fröhlich Hamiltonian, which explicitly accounts for the electron-
phonon (LO phonons at k = 0) interaction (Fröhlich et al. 1950, advanced by Lee
et al. 1953). Alternatively, it can be described with the Feynman Hamiltonian
(Feynman path integral, Feynman 1955), which simulates the virtual phonons by
a fictitious particle that interacts with the electron via a harmonic potential (Peeters
and Devreese 1984).
The size of the polaron is measured by the extent of the lattice distortion caused
by the electron. Large polarons extend substantially beyond nearest-neighbor
distances; small polarons do not. For a review, see Velasco and García-
Moliner (1997).

1.2.1 Large Polarons and Fröhlich Coupling


In the conventional band model, the polarization of Bloch electrons is included by
using the optical dielectric constant, which describes the interaction with the elec-
trons of each lattice atom. In the discussion of polarons, their polarization also
considers the shift in the position of the nuclei: the static dielectric constant is then
used to account for the more intensive shielding of the Coulomb interaction:
      
e2 e2 e2 1 1 1
! ¼   : (4)
8π eopt e0 r el 8π estat e0 r pol 8π e0 r eopt eopt estat

The net difference unaccounted for in the Bloch electron picture describes the
Coulomb energy of the polaron:
Carrier-Transport Equations 7

Fig. 2 Net potential


distribution assigned to a V(r)
polaron, indicating the rpol
polaron radius rpol, below
which the uncertainty relation r
precludes further
extrapolation of the quasi-
Coulomb potential e/(ε*ε0 r)

Epol/e

e2 1 1 1
Epol, Coul ¼ with ¼  : (5)
8π e e 0 r pol e eopt estat

The resulting net Coulomb potential of the polaron is shown in Fig. 2. Within this
potential funnel, the polaron self-energy can be expressed as the 1s quasi-hydrogen
ground-state energy
e 2 m
stat pol
Epol ¼ EqH : (6)
e mn
It is distinguished from the quasi-hydrogen ground-state energy by replacing estat
with e , and the electron effective mass with that of the polaron.
The radius of the polaron can be defined as the corresponding 1s quasi-hydrogen
radius

e m n 4π e e0 ℏ2
r pol ¼ aqH ¼ : (7)
estat mpol mpol e2

When the polaron eigenfunctions are overlapping, the polaron self-energy (Eq. 6)
broadens from a sharp level into a polaron band within which polarons can move
through the crystal. The width of this band can be estimated from the uncertainty by
absorbing or emitting virtual LO phonons as ℏωLO. The corresponding uncertainty
radius of such a polaron is given by the uncertainty distance of finding a particle that
interacted with an LO phonon: it has an energy ℏ2k2/(2mpol) with uncertainty ℏωLO.
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
Therefore, it has an uncertainty in wavenumber of  2mpol ωLO =ℏ, the reciprocal of
which is its corresponding uncertainty in position


r pol ¼ pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi ; (8)
2mpol ℏωLO

which is also used as the radius of a large polaron. For typical compound semi-
conductors, the polaron radius is on the order of several lattice constants.

The Fröhlich Coupling Constant For interaction with LO phonons, it is conve-


nient to express the Fröhlich coupling in terms of the coupling constant αc. It is given
8 K. W. Böer and U. W. Pohl

as the ratio of the Coulomb energy of a polaron, which describes the electron-phonon
interaction, to the energy of the LO phonon, i.e., the predominantly interacting
phonon:
vffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
ffi
u 
u1 e2
u
t2 4π e e0 r pol
αc ¼ : (9)
ℏωLO

Entering the expression for rpol Eq. 7 into Eq. 9 and replacing e* from Eq. 5, we
obtain for the Fröhlich coupling constant αc:
rffiffiffiffiffiffiffiffiffiffiffi  
e2 2 mpol 1 1
αc ¼  ; (10)
8π e0 ℏ ℏωLO eopt estat

which is below unity for good semiconductors. Here αc can be interpreted as twice
the number of virtual phonons surrounding – that is, interacting with – a slowly
moving carrier in the respective band.
Semiconductors without ionic bonding (Si, Ge, α-Sn) have eopt = estat and
consequently αc = 0. Fröhlich coupling constants are therefore much larger for
ionic than for covalent semiconductors and increase with increasing effective charge
and decreasing strength of the lattice binding forces.4 Some values of αc for typical
semiconductors are given in Table 1.

Polaron Energy and Effective Mass Using αc, we can express the polaron energy
as a fraction of the LO phonon energy5:

p2

Epol ¼ Ec  ffi  αc þ 0:01592 α2c þ . . . ℏωLO : (11)


2mpol

Assuming a small perturbation of the parabolic band, Lee et al. (1953) obtain for the
energy dispersion within the band
 
ℏ2 k2 ℏ2 k2 ℏ2 k2
Eð k Þ ¼  αc ℏωLO þ þ . . . ffi αc ℏωLO þ ; (12)
2 mn 12 mn 2 mpol

which yields for the polaron mass

4
A decrease of lattice-bonding forces corresponds also to a decreased Debye temperature; see also
Sect. 1.1.2 of chapter ▶ “Phonon-Induced Thermal Properties”.
5
This result is obtained from Fröhlich et al. (1950) and with a variational method from Lee et al.
(1953). It can be used
up
to αc  1. Earlier results from Pekar, using an adiabatic approximation,
yielded Epol ¼  α2c =3π ℏωLO , which gives a lower self-energy than Eq. 11 in the range of
validity αc < 1.
Carrier-Transport Equations 9

Table 1 Fröhlich coupling constants ac of various solids


Solid αc Solid αc Solid αc Solid αc
InSb 0.02 GaP 0.201 CdTe 0.35 ZnS 0.63
InAs 0.05 GaN 0.48 CdSe 0.46 ZnO 1.19
InP 0.15 AlSb 0.023 CdS 0.51
InN 0.24 AlAs 0.126 CdO 0.74 KI 2.50
GaSb 0.025 AlP 0.49 ZnTe 0.33 KBr 3.05
GaAs 0.068 AlN 0.65 ZnSe 0.43 KCl 3.44

mn
mpol, large ffi : (13)
1  α6c

Equation 13 may be used for small αc, i.e., a weak electron-phonon interaction. For
larger values αc  1 , the following approximation is used for the polaron mass
(Feynman 1955):

16 α4c mn
mpol, large ffi : (14)
81π 4
Materials with large polarons that show a significant increase in the effective mass
are silver halides. These have an intermediate coupling constant (1.60 and 1.91 for
AgBr and AgCl, respectively).
The variational method of Lee et al. (1953) is used to compute intermediate
coupling. Many II-VI and some III-V compounds have αc values that make these
large polarons sufficiently distinct from electrons (see Table 1 and Evrard 1984).
Their effective mass can be determined by cyclotron resonance (Peeters and
Devreese 1984). The Landau levels appearing in a magnetic field are shifted by ΔE
ffi αc ℏωLO for αc 1; when the cyclotron resonance frequency approaches ωLO,
this level splits into two peaks indicating the strength of the electron-lattice interac-
tion. For a more exact approximation using a path integral formulation, see Feynman
(1955). For reviews, see Kartheuser et al. (1979), Bogoliubov and Bogoliubov Jr.
(1986), and Devreese (1984).
In Table 2, the properties of large polarons for a number of crystals are listed. We
note the inverse relation of rpol and Epol; for III-V semiconductors, the polaron mass
is close to the electron rest-mass, while it increases for II-VI semiconductors and gets
large for halides. Correspondingly mn/mpol reflects basically the effective electron
mass in typical semiconductors.
In the presence of a magnetic field, the cyclotron behavior of polarons within the
corresponding Landau levels is observed instead of the bare Bloch band mass of carriers.
For a transition from Landau level n to n + 1, Bajaj (1968) obtained for polarons

  
αc  3 ω2 ℏk2
ωc, pol ¼ ωc 1   αc c þnþ1 ; (15)
6 20 ωLO 2 mn ωc
10 K. W. Böer and U. W. Pohl

Table 2 Polaron Solid rpol (Å) Epol (meV) mpol/m0 mn/mpol


parameters of various solids
InSb 105 0.5 1.00 0.014
InAs 73.9 1.5 1.01 0.023
InP 33.8 5.2 1.02 0.078
GaSb 52.2 0.9 1.01 0.047
GaAs 38.5 2.6 1.01 0.068
GaP 21.1 6.5 1.02 0.175
CdTe 42.8 8.1 1.07 0.107
CdSe 33.1 12.2 1.08 0.14
CdS 25.3 20.2 1.10 0.126
ZnTe 30.5 8.4 1.06 0.169
ZnSe 27.0 13.1 1.08 0.183
ZnS 17.7 30.8 1.14 0.31
ZnO 14.8 64.8 1.18 0.276
KI 25.6 44.8 1.71 0.56
KBr 22.1 64.3 2.03 0.75
KCl 18.2 92.2 2.26 0.97

where ωc is the cyclotron-resonance frequency of electrons (Eqs. 24 and 25 of


chapter ▶ “Bands and Bandgaps in Solids”).

1.2.2 Small Polarons and Criteria for Different Polarons


Small polarons were introduced by Tjablikov (1952) and further analyzed by Hol-
stein (1959). They are observed when the coupling constant is larger than 5. This
strong interaction causes self-trapping: the wavefunction corresponds to a localized
electron; the tight-binding approximation is appropriate for a mathematical descrip-
tion (Tjablikov 1952). Small polarons are distinguished from electrons by a rather
large polaron self-energy Epol. The optical energy necessary to bring a small polaron
into the band, i.e., to free the self-trapped electron, was estimated by Pekar to be

Epol,opt,small ffi 0:14 α2c ℏωLO : (16)

Their thermal ionization energy Epol,th,small is considerably smaller, typically ~(1/3)

Epol,opt,small, because of the strong lattice coupling; see Sects. 1.2.1 and 1.2.2 of chapter
▶ “Optical Properties of Defects”. For further discussion of Epol, see the review of
Devreese (1984).
The effective mass of small polarons is given by (see Appel 1968):

mn α2c
mpol,small ffi : (17)
48
Examples for materials with small polarons are narrow-band semiconductors with
large αc values (~10), e.g., transition-metal oxides such as NiO or the molecular
crystals described in Sect. 3 of chapter ▶ “The Origin of Band Structure” with their
Carrier-Transport Equations 11

particularly small band dispersion (see Sect. 4.3 of chapter ▶ “Quantum Mechanics of
Electrons in Crystals”). Due to this narrow bandwidth, the band conductivity is usually
disturbed by phonon scattering in organic crystals at room temperature, and hopping
mechanisms may prevail even if the material has crystalline structure; see Sects. 4.3
and 5.1.2 of chapter ▶ “Carrier Transport Induced and Controlled by Defects”. A
review for small polarons is given by Emin (1973); for a summary of polaron mobility
due to various scattering mechanisms, see Appel (1968) and Evrard (1984).
When the coupling to the lattice is strong enough so that the polarizing electron
produces a significant Coulomb funnel, a second electron with opposite spin may
be trapped within the same funnel (Chakraverty and Schlenker 1976). This
bipolaron again can move through the lattice at an energy below that of a free
electron and with an effective mass somewhat larger than that of two free
electrons (Böttger and Bryksin 1985). For bipolarons in organic polymers, see
Brazovskii et al. (1998). A significant difference, compared with two free elec-
trons or two independent polarons, is the fact that the new quasiparticle bipolaron
has zero spin and consequently acts as a boson. It is also referred to as a Cooper
pair, and its formation is used to explain superconductivity; see Sect. 1.2 of
chapter ▶ “Superconductivity”.

Existence Criteria for Different Polarons There are three different energies, the
relative magnitudes of which determine the preferred existence of large polarons,
small polarons, or bipolarons. These are6

• The relaxation energy Erelax ¼ αc ℏωLO


• The transfer energy J ¼ ℏ2 =ðncoord m a2 Þ
• The Hubbard correlation energy U ¼ U0  λ2 =β (Sect. 2.8 of chapter ▶ “Deep-
Level Centers”)
Here ncoord is the coordination number, m* is the effective mass, a is the nearest-
neighbor distance, λ is an electron-lattice coupling constant, and β is an elastic
restoring term.
We distinguish:
1. The competition between band formation and lattice relaxation
(a) Erelax J large polaron
(b) Erelax  J small polaron
2. The competition between lattice relaxation and carrier correlation
(a) Erelax U large polarons remain at separate sites.
(b) Erelax  U electrons prefer to share sites.
3. Sign change in carrier-correlation energy
U < 0 bipolaron formation

6
More precisely, these three energies are: Erelax, the energy given to phonons during lattice
relaxation; J, the bandwidth of a band created by free, uncoupled carriers of the given density;
and U, the energy necessary to put two carriers with opposite spin on the same lattice site.
12 K. W. Böer and U. W. Pohl

For further discussion, see Toyozawa (1981).

Electrons or Polarons in Semiconductors In a rigorous presentation, all elec-


trons or holes near the band edges in equilibrium with the lattice should be
replaced by polarons. Since for most semiconductors the difference between
electrons and polarons is very small (Table 2), it is justified to proceed with the
conventional description, using instead Bloch electrons and holes. Even in such
semiconductors, this is no longer sufficient in the neighborhood of lattice defects
with strong electron coupling. Here, the Huang-Rhys factor S, rather than the
Fröhlich coupling constant αc, is used, and the polaron picture is applied in order
to obtain reasonable results for the defect and lattice relaxation in agreement with
the experiment – see Sect. 1.2.2 of chapter ▶ “Optical Properties of Defects”.
Polaron effects are also significant in some amorphous semiconductors (Cohen
et al. 1983).

2 Conductivity and Mobility of Carriers

Carrier transport proceeds under external forces,7 resulting in drift, and under
internal quasiforces, resulting in diffusion. This may involve different charged
particles which contribute additively to the current or have an indirect effect when
it involves neutral particles, e.g., excitons (see Sect. 1.2 of chapter ▶ “Excitons”).
Exciton diffusion plays a major role in devices fabricated from organic semicon-
ductors; for a review see Mikhnenko et al. (2015).
Carrier transport occurs in bands near the band edges, i.e., near Ec for electrons
and near Ev for holes. For materials with a sufficiently large defect density, carrier
transport may proceed also via tunneling between trapping states. It may also involve
carriers hopping from traps into the band or hopping of self-trapped small polarons.
Trapped carriers travel a short distance in the band and later are recaptured, then
reemitted, and so on; an analogous process occurs with small polarons – see chapter
▶ “Carrier Transport Induced and Controlled by Defects”.
All of these processes add up to produce the total current and usually have vastly
different magnitudes. Ordinarily, only one transport process predominates in homoge-
neous semiconductors. In nonhomogeneous materials, however, at least two and fre-
quently four contributions are important in different regions of the devices. These are
drift and diffusion currents of electrons and holes. First, a rather simple picture of the
carrier transport is presented, which serves as guidance for a more sophisticated
approach in later chapters.

7
Strictly speaking, steady-state carrier transport is due to external forces only. The diffusion current
originates from a deformed density profile due to external forces and is a portion of the conven-
tionally considered diffusion component. The major part of the diffusion is used to compensate the
built-in field and has no part in the actual carrier transport: both drift and diffusion cancel each other
and are caused by an artificial model consideration – see Sect. 3.4.
Carrier-Transport Equations 13

Fig. 3 Classical velocity vmp v


distribution, with root mean av
1 vrms
square, average, and most
probable velocities identified

vn /vmp
0.5

0
0 50 100 150 200 250
vn (km/s)

At finite temperatures, carriers are found above the edge of the respective band
according to their statistical distribution function. In semiconductors, they usually
follow the Boltzmann distribution function within the bands when they are not
degenerate, i.e., when the carrier densities are below 0.1Nc or 0.1Nv (Eqs. 18 and
22 of chapter ▶ “Equilibrium Statistics of Carriers”). Their thermal velocity, the root
mean square velocity,8 is obtained from the equipartition principle, i.e., kinetic
energy = ½kT per degree of freedom:

hence mn 2 3
v ¼ k T; (18)
2 2
pffiffiffiffiffiffiffiffi 3k T
hv2 i ¼ vrms ¼ ; (19)
mn
or
sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
m0 T ðKÞ
vrms ¼ 1:18
107 ðcm=sÞ: (20)
mn 300

The velocity distribution is illustrated in Fig. 3. In thermal equilibrium and in an


isotropic lattice, the motion of the carriers is random.
The quantum-mechanical model of a periodic potential teaches that, in contrast to a
classical model, an ideal lattice is transparent for electrons or holes within their
respective bands. That is, the carriers belong to the entire semiconductor and as
waves are not localized: their position cannot be identified, except stating that
n carriers (per cm3) are within the given crystal. There is no scattering of carriers
within such an ideal crystal. This behavior of carriers is unexpected in a classical model,

8
The rms (root mean square) velocity vrms, which is commonly used, should be distinguished from
the slightly different average velocity vav ¼ hjvji and from the most probable velocity vmp. Their
pffiffiffiffiffiffiffiffi pffiffiffiffiffiffiffiffi
ratios are vrms : vav : vmp ¼ 3=2 : 4=π : 1 ¼ 1:2247 : 1:1284 : 1, as long as the carriers follow
Boltzmann statistics. For a distinction between these different velocities, see Fig. 3.
14 K. W. Böer and U. W. Pohl

Fig. 4 Random walk of a A


carrier with (red curves) and
without (blue lines) an B
external electric field. The C
green arrow from B to
C indicates the relative
displacement in field direction
after the indicated nine
scattering events

which visualizes the filling of space with atomic spheres and expects only very limited
possibilities for an electron traversing between these spheres without being scattered.
The introduction of phonons and crystal defects provides centers for scattering. In
this way, a carrier motion results, which can be described as a Brownian motion9 with a
mean free path commensurate with the average distance between scattering centers.
This distance is several hundred angstroms in typical crystalline semiconductors, i.e.,
the mean free path extends to distances much longer than the interatomic spacing.
When a carrier responds to an external field, it is accelerated in the direction of the
electric field. Many important features of the carrier motion can be explained by
assuming only inelastic scattering. Since it takes place at defects, which themselves
are in thermal equilibrium with the lattice, the carrier tends to lose the excess energy
gained between the scattering events. Figure 4 illustrates the typical motion of a
carrier with and without an external electric field, assuming for both cases identical
scattering events. The changes due to the field are exaggerated; under normal
external fields, the changes from the random walk without field are very small
perturbations barely being visible in the scale of Fig. 4.

Sign Conventions In previous chapters, the elementary charge is used as e ¼ jej.


When the transport of electrons and holes is discussed, it is instructive to discuss the
proper signs: e for electrons and +e for holes. This has an influence on derived
parameters, e.g., the mobility, as will be discussed in Sect. 2.2.
The electric field10 F is defined as the negative gradient of the vacuum level. The
bias V is conventionally labeled + for the anode and – for the cathode, while the
electrostatic potential ψ has the opposite signs: electrons have a larger potential
energy at the cathode than at the anode. The relation to the field F is therefore

dV dψ
F¼ ¼ : (21)
dx dx
With Ec ¼ jej ψ þ const; it yields a positive field when the band slopes downward
from the cathode toward the anode, giving the visual impression that electrons “roll

9
This motion resembles a random walk (Chandrasekhar 1943).
In chapters dealing with carrier transport, F is chosen for the field, since E is used for the energy.
10
Carrier-Transport Equations 15

downhill” and holes “bubble up.” We will use the proper signs in the following
sections, however, reverting back to the commonly used e ¼ jej later in order to
avoid confusion in comparison to familiar descriptions.
When expressing forces, we need to distinguish the sign of the carrier; therefore,
for an accelerating force we have

Force ¼ ðe FÞn ¼ ðþe FÞp ; (22)

with subscripts n and p for electrons and holes.

2.1 Electronic Conductivity

In following the arguments introduced by Drude (1900; later refined by Lorentz


1909, and Sommerfeld 1928), electrons are accelerated in an electric field F by the
force eF:

dv
mn ¼ e F: (23)
dt
During a free path, the electron gains an incremental velocity, for an arbitrarily
chosen field in x direction F = (Fx, 0, 0):

e
Δvx ¼  Fx τsc : (24)
mn
After averaging the incremental velocity between collisions and replacing the scat-
tering time τsc with the average time τ between scattering events, we obtain the drift
velocity vd:
e
vd ¼  τ Fx : (25)
mn
With an electron density n and a charge e, we obtain for the current density for
electrons

j n ¼ e n vd ; (26)

or, introducing the electron conductivity σ n,

e2
jn ¼ σ n F, with σ n ¼ τn n : (27)
mn
In a homogeneous semiconductor, the external field is given by the bias V, divided
by the electrode distance d, yielding Ohm’s law
16 K. W. Böer and U. W. Pohl

V V
jn ¼ σ n F ¼ σ n ¼ ; (28)
d AR

with A as the area of the semiconductor normal to the current and the resistance
ρn d d
R¼ ¼ ; (29)
A A σn
where ρn is the specific resistivity ρn ¼ 1=σ n :

Joule’s Heating The additional energy gained by the accelerated electrons from the
external electric field is delivered to the lattice during inelastic collisions, generating
phonons. This Joule’s heating is given by the power density p which n electrons
transfer to the lattice while travelling at the average constant drift velocity vd despite
the accelerating force – eF; applying the sign convention used above we obtain

j2n
p ¼ n e vd F ¼ jn F ¼ σ n F2 ¼ : (30)
σn

Bloch Oscillations When a Bloch electron is accelerated in an external field, it moves


up relative to the conduction-band edge. When the band is narrow and the field is high
enough, the electron can cross the center of the conduction band; as it approaches the
upper band edge before scattering occurs, its effective mass becomes negative (Sect. 2.2
of chapter ▶ “The Origin of Band Structure”). It is consequently decelerated and finally
reflected at the upper conduction-band edge. It hence moves down in the conduction
band until the process repeats, and the Bloch electron undergoes an oscillatory motion,
the Bloch oscillation. This was proposed early (Esaki and Tsu 1970), but can hardly be
observed in bulk semiconductors, because conduction bands are usually too wide for an
undisturbed motion high into the band before scattering occurs. In superlattices,
however, minibands are narrow enough to permit the observation of Bloch oscillations;
see, e.g., Shah (1994), Leo (1996), and Waschke et al. (1994).

2.2 Electron Mobility

The quantity

e
τ n ¼ μn (31)
mn

is the electron mobility, since carriers are more mobile when they experience
less scattering, i.e., the time between collisions is larger, and when their effective
mass is smaller, i.e., they can be accelerated more easily. With e for electrons and
+e for holes, the mobility is negative for electrons and positive for holes, while
Carrier-Transport Equations 17

the conductivity (/ e2, see Eq. 27) is always positive. Conventionally, however,
μn ¼ jμn j is used, and we will follow this convention here.
The electron conductivity and hole conductivity are given by

σ n ¼ e μn n and σ p ¼ e μp p: (32)

Gas-Kinetic Model for Electron Scattering Different types of lattice defects are
effective to a differing degree in carrier scattering. In a simple gas-kinetic model,
scattering centers have a well-defined scattering cross-section sn; a scattering
event, i.e., a marked deflection11 from an otherwise straight carrier path with
exchange of momentum and/or energy, takes place when the carrier approaches
the scattering center within its cross section. A mean free path λn can then be
derived by constructing a cylinder of cross section sn around an arbitrary straight
carrier path and computing the average distance from the last scattering center to
which this cylinder will extend until it incorporates the centerpoint of the next
scattering center. At this length the cylinder volume λn sn equals the average
volume 1/Nsc (cm3) that one of these centers occupies, where Nsc is the density
of scattering centers; hence

1
λn ¼ : (33)
sn N sc
Consequently, the time between scattering events is given by

1
τsc ¼ : (34)
vrms sn N sc
This time is used to obtain an estimate for the carrier mobility in the Drude
approximation:

e e λ
μn ¼ τsc ¼ ; (35)
mn mn vrms

or
m0

μn ¼ 1:8
1015 τsc ðsÞ cm2 =Vs ; (36)
mn
or, using the expression of Eq. 19 for vrms:

Often a minimum scattering angle of 90 is used to distinguish scattering events with loss of
11

memory from forward scattering events – see Sect. 4.6.1.


18 K. W. Böer and U. W. Pohl

a 5 b5

mn /m0= 0.1 mn /m0= 0.1


4 4
log (μ/(cm2/Vs))

0.4 0.4 1
1

3 0.7 3 0.7

2 2

1 1
–14 –13 –12 –11 –10 1 2 3 4 5
log (τsc/s) log (λ/Å)

Fig. 5 Carrier mobility as a function of (a) the time between scattering events τsc and of (b) the
mean free path λ, with the effective electron mass mn as family parameter

 3=2  
m0 300 1=2 2

μn ¼ 1:5
λðÅÞ cm =Vs : (37)
mn T ðK Þ

The dependence of the carrier mobility on the mean scattering time and mean free
path is shown in Fig. 5. The application of this simple gas-kinetic model, however,
has to be taken with caution because of its simplified assumptions. Generally, it
yields too large densities of tolerable scattering centers.

3 Currents and Electric Fields

3.1 Drift Current in an Electric Field

The drift current is the product of the elementary charge, the carrier mobility, the
single carrier density (derived in chapter ▶ “Equilibrium Statistics of Carriers”), and
the electric field. For electrons or holes, it is

jn, drift ¼ e n μn F or jp, drift ¼ e p μp F: (38)

External Electric Field In homogeneous semiconductors, disregarding space-


charge effects near interfaces, and for steady-state conditions – assumed with few
exceptions throughout this book – the electric field is given by the applied voltage
(bias) divided by the distance between the electrodes in a one-dimensional geome-
try; see Fig. 6:
Carrier-Transport Equations 19

Fig. 6 Preferred quasi-one-


dimensional geometry with V
band diagram subject to an
external bias V, resulting in
band tilting

E d

Ec
eV
EF

Ev

V
F¼ : (39)
d
For the field concept to apply, the distance d between electrodes must also be
large compared to the interatomic spacing. The field can then be expressed by the
macroscopic sloping of the bands12:
1 d Ec 1 d Ev
F¼ ¼ : (40)
e dx e dx
It is also given, and more importantly so, by the slope of the Fermi potential (see Sect.
3.3) which, within the homogeneous material, is the same as the slope of the bands:
1 d EF
F¼ : (41)
e dx
The bias is expressed as the difference of the Fermi levels between both electrodes;
compare Fig. 6.
When using the electrostatic potential ψ n with
d ψn
 ¼ F; (42)
dx
the drift current can be expressed as a product of the electrical conductivity and the
negative gradient of this potential:

12
As a reminder: here and in all following sections, |e| is used when not explicitly stated differently.
20 K. W. Böer and U. W. Pohl

Fig. 7 Band diagram for a E


ZnSeuS1-u mixed crystal with
linearly varying composition Ec
u along the x axis. At x =
0, the material is ZnSe; at x = EF
d, the material is ZnS

Ev

ZnSe ZnSeuS1-u ZnS

u
1 0

0 x d

d ψn d ψp
jn,drift ¼ σ n or jp,drift ¼ σ p : (43)
dx dx
For reasons to become apparent below, two electrostatic potentials
are introduced:

ψn
and ψ p for conduction and valence bands, respectively, with e ψ n  ψ p ¼ Eg .
In a homogeneous semiconductor in steady state and with vanishing space
charge, these drift currents are the total currents, and the slopes of both potentials
are the same. There are special cases, however, in which the band edges of the
valence and conduction bands are no longer parallel to each other. One of these will
be mentioned briefly in the following section.

The Built-In Electric Field A semiconductor with a graded composition produces


a position-dependent, varying bandgap energy. If this composition varies smoothly
without steps, one or both bands are sloped without an applied bias, representing
built-in fields – see Sect. 3.4. As an example, in ZnSeuS1-u, there is complete
miscibility in the entire range (0 u 1), with the S-Se sublattice being a statistical
alloy.13 The bandgap energy changes linearly14 from 2.45 eV for ZnSe at the left side
of the crystal shown in Fig. 7 to 3.6 eV for ZnS at its right side – see also Fig. 24 of

13
The atoms in the alloyed lattice (or here in the anion sublattice) are statistically arranged. Strictly
speaking, this causes random fluctuation of the composition in a microscopic volume element of the
crystal and results in a local fluctuation of the bandgap energy due to a locally varying parameter x;
as a result, extended or localized states with energies Ec,v(x) close to the band edge Ec, v ðxÞ of the
mean composition x are formed, leading to some tailing of the band edges (Sect. 3.1 of chapter
▶ “Optical Properties of Defects”). The composition parameter x used in the text actually refers to
the mean composition, and the effect of band tailing is neglected here.
14
Major deviations from linearity of Eg with composition are observed when the conduction-band
minimum lies at a different point in the Brillouin zone for the two end members. One example is the
alloy of Ge and Si. Other deviations (bowing – see Sect. 2.1 of chapter ▶ “Bands and Bandgaps in
Solids”) are observed when the alloying atoms are of substantially different size and electronegativity.
Carrier-Transport Equations 21

chapter ▶ “Bands and Bandgaps in Solids”. ZnSe and ZnS are both n-type materials.
Depending on doping, the Fermi level in ZnSe can be shifted easily between 0.8 and
0.2 e V below Ec and in ZnS between 1.0 and 0.4 eV. Depending on the doping profile
in the mixed-composition region, a wide variety of relative slopes, including non-
monotonic slopes, of valence and conduction bands can be designed for a
vanishing bias, i.e., for a horizontal Fermi energy EF. In Fig. 7, an example with
opposite and linear sloping of Ec(x) and Ev(x) is shown, resulting effectively in a
built-in field (see Sect. 3.4) of opposite sign for electrons and holes. In thermo-
dynamic equilibrium, however, there is no net current in spite of the sloping
bands. This is accomplished by exact compensation of finite drift currents with
opposing diffusion currents, which self-consistently determine the slopes of the
bands. An example for the application of such graded AluGa1-uAs composition is
given in Horio et al. (1999).
The change in the bandgap energy can be expressed as

Ec ðxÞ ¼ Ev ðxÞ þ Eg0 þ ΔEg ðxÞ; (44)

where x is a spatial coordinate. Using a conventional asymmetry factor AE, which


measures the fraction of the bandgap change ΔEg(x) occurring in the conduction
band relative to the horizontal Fermi level, Eq. 44 is also written

Ec ðxÞ ¼ Ec ðx ¼ 0Þ þ AE ΔEg ðxÞ ¼ eψ n ðxÞ (45)

and

Ev ðxÞ ¼ Ev ðx ¼ 0Þ  ð1  AE Þ ΔEg ðxÞ ¼ eψ p ðxÞ: (46)

The corresponding built-in fields for electrons and holes are given by

@ΔEg ðxÞ @ψ
Fn ¼ AE ¼ n (47)
e@x @x
and

@ΔEg ðxÞ @ψ p
Fp ¼  ð AE  1 Þ ¼ ; (48)
e@x @x
justifying the introduction of separate electrostatic potentials for electrons and holes:
with ΔEg ðxÞ 6¼ 0 and AE 6¼ 1/2, we have Fn 6¼ Fp.

3.2 Diffusion Currents and Total Currents

Carrier diffusion by itself can be observed when the external field vanishes and a
concentration gradient exists. An example in which these conditions are approxi-
mately fulfilled is the diffusion of minority carriers created by an inhomogeneous
22 K. W. Böer and U. W. Pohl

Fig. 8 Illustration of the n


derivation of the diffusion
current in a medium with
inhomogeneous carrier
density n(x). Arrows indicate n0+dn/2
diffusion currents n0
n0-dn/2

x0-dx/2 x0 x0+dx/2 x

optical excitation (see, e.g., Najafi et al. 2015). The diffusion current is proportional
to the diffusion coefficient D and to the carrier-density gradient; for electrons or
holes, it is

dn dp
jn,diff ¼ e Dn or jp,diff ¼ e Dp : (49)
dx dx

The negative sign of the hole current is due to the fact that in both equations e
¼ jej is used. The diffusion current can be derived as the difference between two
currents caused by a completely random motion of carriers originating in adjacent
slabs with slightly different carrier densities (Fig. 8). The current originating at x0 +
dx/2 and crossing the interface at x0 from right to left, is caused by the Brownian
motion of electrons of a density n0 + dn/2. It is given by
 
dn v2rms τn
j n,diff ¼ e n0 þ : (50)
2 3 dx

The current crossing the boundary from left to right is accordingly given by
 
! dn v2rms τn
j n, diff ¼ e n0  : (51)
2 3 dx

The current is proportional to the carrier velocity vrms and the carrier mean free
path λn. In turn, λn is given by vrmsτn. The factor 13 arises from gas-kinetic arguments
when the root mean square velocity is obtained from an isotropic velocity distribu-
tion: v2 ¼ v2rms ¼ v2x þ v2y þ v2z ; with v2x ¼ v2y ¼ v2z , we obtain the relation v2x ¼ 13v2rms
for the x-component used in Eq. 50.
The difference of both currents (Eqs. 50 and 51) is the net diffusion current
Carrier-Transport Equations 23

! v2rms τn dn
jn, diff ¼ j n, diff  j n, diff ¼ e ; (52)
3 dx
with the diffusion coefficient given by

v2rms τn
Dn ¼ : (53)
3
By using v2rms ¼ 3k T=mn (Eq. 19), we obtain the more commonly used equation
for the diffusion current

dn dp
jn,diff ¼ μn k T and jp,diff ¼ μp k T : (54)
dx dx

Both diffusion currents for electrons and holes have the same negative sign for a
positive gradient of n(x) or p(x) when recognizing that μn is negative and μp is
positive. However, since the conventional notation with μn ¼ jμn j is used, the
difference in signs appears.

Maximum Diffusion Currents As the gradient of the carrier density increases, the
diffusion current increases proportionally to it (Eq. 49). However, this proportion-
ality is limited, when the density gradient becomes so steep that the reverse current
(Eq. 51) becomes negligible compared to the forward current (Eq. 50).
When increasing the distance dx to the mean free path λn, we obtain from Eq. 52
!
with j n, diff j n, diff for the maximum possible diffusion current through a planar
surface

v2rms τn
jn,diff ,max ¼ e n0 ; (55)
3 λn
or, for carriers following Boltzmann statistics and within a device with planar
geometry,

en
jn,diff ,max ¼ pffiffiffiffiffi vrms : (56)

This current is known as the Richardson-Dushman current (Dushman 1930). It


is equal to the thermionic emission current into the vacuum if the semiconductor
is cut open at x0 (Fig. 8) and if a vanishing work function is assumed – that is, if
all electrons in the conduction band at x0, with a velocity component toward the
surface, could exit into the vacuum (see also Simoen et al. 1998).
24 K. W. Böer and U. W. Pohl

Fig. 9 Electron density and


electrostatic potential
n
distribution in the Boltzmann
region in thermal equilibrium
(schematic)
n0+Δn
n0

x0 x0+Δx x
V

V0
ΔV=FΔx

x0 x0+Δx x

The Einstein Relation Comparing the diffusion equations, Eq. 49 with Eq. 54, we
obtain a relation between the diffusion constant and the carrier mobility:

μn, p k T
Dn, p ¼ ; (57)
e

which is known as the Einstein relation and holds for systems that follow Boltzmann
statistics. This can be seen from the following arguments. In thermal equilibrium, the
total current, as well as each carrier current, vanishes: j jn jp 0. The electron
current is composed of drift and diffusion currents – see Eq. 61; hence,

dn
μn nF þ Dn ¼ 0; (58)
dx
which can be integrated to yield

nðx0 þ ΔxÞ ¼ nðx0 Þexpðμn FΔx=Dn Þ: (59)

On the other hand, electrons obey the Boltzmann distribution in equilibrium in


the conduction band of a semiconductor. Their surplus energy, obtained in an electric
field at a distance Δx, is ΔE = eΔV = e FΔx (Fig. 9), yielding a density
Carrier-Transport Equations 25

 
eF Δx
nðx0 þ ΔxÞ ¼ nðx0 Þexp  : (60)
kT

A comparison of the exponents in Eqs. 59 and 60 yields the Einstein relation


(Eq. 57). In case the Boltzmann distribution is not fulfilled (degeneracy), a
generalization of Eq. 57 can be derived involving Fermi integrals (Landsberg
1952).
The assumptions used beyond the Boltzmann distribution are that of a one-carrier
model near equilibrium and that the total current is small compared to drift and
diffusion currents; hence, Eq. 58 holds. At high fields, one or more of these
conditions are no longer fulfilled. Consequently, the Einstein relation needs to be
modified – see Sect. 1 of chapter ▶ “Carrier Scattering at High Electric Fields” and
Kan et al. (1991). For nonparabolic bands, see Landsberg and Cheng (1985). For
hot-carrier diffusion at low temperature in quantum wells, see Chattopadhyay
et al. (1989).

Total Currents The total current is given as the sum of drift and diffusion currents.
For electrons, we have

dn
jn ¼ jn,drift þ jn,diff ¼ e μn n F þ Dn ; (61)
dx

and for holes


dp
jp ¼ jp,drift þ jp,diff ¼ e μp p F  Dp : (62)
dx
The total carrier current is the sum of both,

j ¼ jn þ jp : (63)

In homogeneous semiconductors, only one of the four components is usually


predominant, while in a pn junction with sufficient bias, each one becomes predom-
inant within a different region.

3.3 Electrochemical Fields and Quasi-Fermi Levels

For evaluating the total current, we need to consider the gradient of ψ(x) for the drift
component and the gradient of n(x) for the diffusion component. It is instructive to
deduce the electron and hole current for thermal equilibrium. In thermal equilibrium,
n is given by the Fermi distribution (Eq. 10 of chapter ▶ “Equilibrium Statistics of
Carriers”). When the Fermi energy is separated by several kT from the band edge, we
can disregard the 1 in the denominator of Eq. 10 in chapter ▶ “Equilibrium Statistics
of Carriers” and approximate this equation with the Boltzmann distribution, yielding
26 K. W. Böer and U. W. Pohl

Nc
Ec ðxÞ  EF ðxÞ ¼ k T ln : (64)
nð x Þ

Replacing Ec(x) with eψ ðxÞ þ c and differentiating both sides of Eq. 64 with
respect to x, we obtain after division by e

dψ n 1 dEF k T 1 dn
  ¼ : (65)
dx e dx e n dx
After multiplying both sides with σ n ¼ e μn n and rearranging, we obtain

1 dEF dψ dn
σn ¼ σ n n þ μn k T : (66)
e dx dx dx
The right-hand side is the total electron current; thus, the left-hand side must also be
equal to jn:

1 dEF
jn ¼ σ n : (67)
e dx
Since in thermal equilibrium, i.e., with vanishing external field, the Fermi level must
be constant (horizontal in an E(x) presentation), we conclude that the electron and
hole current must vanish separately in equilibrium:

dEF
0 ) jn jp 0: (68)
dx

Quasi-Fermi Levels In steady state, e.g., with a constant external excitation (chap-
ter ▶ “Carrier Generation”), the electron and hole densities deviate from thermody-
namic equilibrium values. Nevertheless, we may use the Fermi distribution to
describe their density in the bands, using the quasi-Fermi levels EFn and EFp
according to the definition equations

1
n Nc   ; (69)
Ec  EFn
exp þ1
kT

and

1
p Nv   ; (70)
EFp  Ev
exp þ1
kT

with EFn 6¼ EFp. This rather useful approximation introduces errors which may or
may not be acceptable depending on the cause for deviation from the thermal
equilibrium. In general, the error is quite small for optical excitation and for low
Carrier-Transport Equations 27

Fig. 10 Band model with E


external excitation resulting in
Ec
a split of the Fermi level EF
into two quasi-Fermi levels EFn
EFn and EFp. The scheme EF
represents the zero-field case

EFp

Ev

external fields. For high external fields, the distribution function is substantially
deformed (see chapter ▶ “Carrier Scattering at High Electric Fields”), and a more
sophisticated approximation is required.
With optical or field-induced carrier generation discussed in chapter ▶ “Carrier
Generation”, n and p are increased above their thermodynamic equilibrium value;
hence, EFp < EF < EFn, resulting in a decreased distance of both quasi-Fermi levels
from their corresponding bands; see Fig. 10. In certain cases, the recombination may
be increased above the equilibrium value, as, for instance, in a pn junction in reverse
bias; here, EFn can drop below EFp.
In using the same algebraic procedure as described in the previous section, we
have for the total electron current in steady state:

1 dEFn
jn ¼ σ n ; (71)
e dx
this means, the total electron current is proportional to the negative slope of the
quasi-Fermi potential, as the drift current is proportional to the negative slope of the
electrostatic potential (Eq. 43). For both currents, the conductivity is the proportion-
ality constant.
In order to emphasize this similarity, we define the electrochemical potentials for
electrons and holes:

1 1
φn ¼ EFn and φp ¼ EFp : (72)
e e
The total currents can now be expressed as

@φn @φp
jn ¼ σ n and jp ¼ σ p : (73)
@x @x
For homogeneous semiconductors with homogeneous generation of carriers, these
currents become the drift currents, and Eq. 73 becomes equal to Eq. 43.
In steady state, the total current is divergence-free, i.e., jn+ jp = const (for
included carrier generation see chapter ▶ “Photoconductivity”). Therefore,
28 K. W. Böer and U. W. Pohl

@φn @φp
σn þ σp const (74)
@x @x
or
@φn @φp
μ n nð x Þ þ μ p pð x Þ const: (75)
@x @x
Since a semiconductor is predominantly either n- or p-type, except for the inner part
of a junction, we usually can neglect one part of the sum. For example, for the n-type
region,

@φn
nð x Þ const; (76)
@x
i.e., if there is a gradient in the carrier density, then the highest slope in φn(x) is
expected where the carrier density is lowest for an inhomogeneous n(x) distribution.
From Eq. 75, we also conclude that for vanishing currents in steady state, the
slopes of the quasi-Fermi potentials must be opposite to each other. The lower the
corresponding carrier densities, the higher the slopes:

@φn @φp
μ n nð x Þ ¼ μp pðxÞ : (77)
@x @x

Summary: Potential Gradients and Currents The various currents in a semicon-


ductor can be expressed in a similar fashion. They are proportional to the negative
gradient of electrostatic or electrochemical potentials with the conductivity as a
proportionality factor:

@φn @φp
jn ¼ σ n , jp ¼ σ p ;
@x @x
dψ n dψ p
jn, drift ¼ σ n , jp, drift ¼ σ p ; (78)
dx dx

ð φn  ψ n Þ φp  ψ p
jn, diff ¼ σ n , jp, diff ¼ σ p :
@x @x

3.4 Carrier Distributions in External and Built-In Fields

An external bias resulting in a surface charge on the two electrodes leads to an


external field with no space charge within the semiconductor (Fig. 6). However, if
Carrier-Transport Equations 29

there are inhomogeneities in the distribution of charged donors or acceptors15 or


spatially inhomogeneous compositions in an alloy semiconductor (Fig. 7), regions
with a space-charge density ρ exist within the semiconductor. This charge density
causes the development of an internal field Fint according to the Poisson equation
dFint ρ
¼ : (79)
dx e e0
The acting resulting field is the sum of both internal and external fields,

F ¼ Fint þ Fext : (80)

An external and internal field of equal magnitude results in the same slope of the
bands. Therefore, this distinction between internal and external fields is usually not
made, and the subscripts at the fields are omitted. There are, however, drawbacks in
such a general description of fields, which can best be seen from carrier heating in an
electric field. Carrier heating is used to describe the field dependence of the mobility
(see Sects. 2 and 3 of chapter ▶ “Carrier Scattering at High Electric Fields”) in a
microscopic model: accelerated carriers are shifted up to higher energies within a
band; consequently, their effective mass and the scattering probability change. Usually
the effective mass increases and the creation of phonons becomes easier. The mobility
hence becomes field dependent and usually decreases with increasing field.
Such carrier heating is absent in thermal equilibrium: the carrier gas and the
lattice with its phonon spectrum are in equilibrium within each volume element;
thus, carrier and lattice temperatures remain the same (Stratton 1969). No energy can
be extracted from an internal field, i.e., from a sloped band, due to a space charge in
equilibrium.16 This situation may be illustrated with an example replacing electrical
with gravitational forces: a sloping band due to a space-charge region looks much
like a mountain introduced on top of a sea-level plane, the Fermi level being
equivalent to the sea level. As the introduction of the mountain does little to the
distribution of molecules in air, the introduction of a sloping band does little to the
distribution of electrons in the conduction band. Since there are fewer molecules
above the mountain, the air pressure is reduced, just as there are fewer electrons in a
band where it has a larger distance from the Fermi level as illustrated in Fig. 11.
However, when one wants to conveniently integrate overall altitudes (energies) in
order to arrive at a single number – the air pressure (or electron density) – one must
consider additional model consequences to prevent winds from blowing from the
valleys with high pressure to the mountain top with low pressure by following only
the pressure gradient. Neither should one expect a current of electrons from the

15
pn junctions are the best studied intentional space-charge regions. Inhomogeneous doping
distributions – especially near surfaces, contacts, or other crystal inhomogeneities – are often
unintentional and hard to eliminate.
16
This argument no longer holds with a bias, which will modify the space charge; partial heating
occurs, proportional to the fraction of external field. This heating can be related to the tilting of the
quasi-Fermi levels.
30 K. W. Böer and U. W. Pohl

Fig. 11 Fermi distribution


for different positions in a E
semiconductor at zero-applied Ec
bias with a built-in field region
due to a junction or a gradient
in composition (After Böer
1985a)

EF

1/2
1 x
f (E)

regions of a semiconductor with the conduction band close to the Fermi level, which
results in a high electron density, to a region with low electron density in the absence
of an external field. To prevent such currents in the electron-density model, one uses
the internal fields, i.e., the built-in fields, and balances the diffusion current with an
exactly compensating drift current. The advantage of this approach is the use of a
simple carrier density and a simple transport equation. The penalty is the need for
some careful definitions of transport parameters, e.g., the mobility, when comparing
external with built-in fields and evaluating the ensuing drift and diffusion currents
when the external fields are strong enough to cause carrier heating.

Carrier Concentrations in Built-In or External Fields The carrier distribution


and mobility are different in built-in or external fields. The carrier distribution is
determined relative to the Fermi level. For vanishing bias, the distribution does not
depend on the position; the Fermi level is constant (horizontal). The distribution
remains unchanged when a junction with its built-in field is introduced.17 The
sloping bands cut out varying amounts from the lower part of the distribution,
much like a mountain displaces its volume of air molecules at lower altitudes; see
Fig. 11. The carrier concentration n becomes space dependent through the space

17
With bias, the Fermi level in a junction is split into two quasi-Fermi levels which are tilted,
however, with space-dependent slope. Regions of high slope within the junction region will become
preferentially heated. The formation of such regions depends on the change of the carrier distribu-
tion with bias and its contribution to the electrochemical potential (the quasi-Fermi level). Integra-
tion of transport, Poisson, and continuity equations yields a quantitative description of this behavior
(Böer 1985b).
Carrier-Transport Equations 31

a b
E E

E1

ħω

Ec

EF EF

x x1 x2 x

Fig. 12 Sloping band due to (a) an internal (built-in) field with horizontal Fermi level EF and (b)
due to an external field with parallel sloping of both, bands and Fermi level. The electron
distribution is indicated by a dot distribution, and the action of field and scattering by arrows
(After Böer (1985a))

dependence of the lower integration boundary, while the energy distribution of the
carrier n(E) remains independent in space:
ð1
nð x Þ ¼ nðEÞ dE: (81)
Ec ðxÞ

This is similar to the velocity distribution of air molecules, which is the same at any
given altitude, whether over a mountain or an adjacent plane; whereas the integrated
number, i.e., the air pressure near the surface of the sloping terrain, is not. This does
not cause any macroscopic air motion, since at any stratum of constant altitude, the
molecular distribution is the same; hence, the molecular motion remains totally
random.
In a similar fashion, electrons at the same distance above the Fermi level are
surrounded by strata of constant electron density; within such strata their motion
must remain random. During scattering in thermal equilibrium, the same amounts of
phonons are generated as are absorbed by electrons, except for statistical fluctua-
tions: on the average, all events are randomized. Electron and hole currents both
vanish in equilibrium for every volume element. Figure 12a illustrates such a
behavior.
In an external field, however, bands and Fermi level are tilted parallel to each
other; this means, with applied bias, the carrier distribution becomes a function of the
spatial coordinate as illustrated in Fig. 12b. When electrons are accelerated in the
field, they move from a region of higher density nðE1  EF Þx1 to a region of lower
density nðE1  EF Þx2. These electrons can dissipate their net additional energy to the
lattice by emitting phonons and causing lattice (Joule’s) heating. In addition, while in
32 K. W. Böer and U. W. Pohl

net motion, electrons fill higher states of the energy distribution, thereby causing the
carrier temperature to increase. The carrier motion in an external field is therefore no
longer random; it has a finite component in field direction: the drift velocity vd =
μFext and the collisions with lattice defects are at least partially inelastic. A net
current and lattice heating result.

Field Dependence of Mobilities At higher fields the carrier mobility becomes


field dependent. The difference between the built-in and the external fields relates
to the influence of carrier heating on the mobility, since the averaging process for
determining the mobility uses the corresponding distribution functions. For
instance, with an electric field in the x direction, one obtains for the drift velocity
of electrons
ð
vx f ðvÞ gðvÞ d 3 v
v d ¼ μ n Fx ¼ v x ¼ ð ; (82)
f ðvÞ gðvÞ d 3 v

where g(v) is the density of states in the conduction band per unit volume of velocity
space and d3v is the appropriate volume element in velocity space. If Fx is the built-in
field Fint, then the distribution function is the Boltzmann function fB(v). If Fx is the
external field Fext, the distribution function is modified due to carrier heating
according to the field strength f Fext ðvÞ – see Sects. 4.4 and 4.6. The averaging process
involves the distribution function, which is modified by both scattering and
effective-mass contributions. For a review, see Nag (1980); see also Seeger (1973)
and Conwell (1967). In contrast, when only a built-in field is present, the averaging
must be done with the undeformed Boltzmann distribution, since lattice and electron
temperatures remain the same at each point of the semiconductor.
A more detailed discussion of differences in carrier transport for external, built-in,
and mixed fields is postponed to chapter ▶ “Carrier Scattering at High Electric
Fields” after an explicit introduction of carrier heating.

4 The Boltzmann Equation

There are several simplifying assumptions in the Drude-Sommerfeld approach


introduced in Sect. 2.1 that are not generally valid.
First, most of the scattering events are not inelastic. A carrier usually accumulates
energy during several mean free paths. Each path is interrupted by a mostly elastic
collision until it dissipates the increased energy in one inelastic collision; then it
continues to accumulate energy, and so on. Different kinds of collisions must be
distinguished. This requires replacing the average time between collisions with a
relaxation time, which is typical for the decay of a perturbation introduced by an
applied external force, e.g., an external electric field.
Carrier-Transport Equations 33

Second, the interaction with a scattering center often depends on the energy of the
electron, e.g., the scattering cross-section of ions is energy dependent, and only more
energetic electrons can be scattered inelastically. Therefore, the assumption of a
constant, energy-independent time between scattering events needs to be refined. In
this chapter, we will introduce the basic Boltzmann equation which permits an
analysis of a more advanced description of carrier scattering.

4.1 The Boltzmann Equation for Electrons

A formalism that permits a refinement of the carrier-transport analysis must account


for the change in the population of carriers in space and energy or momentum when
exposed to external forces. This population is described by a distribution function: in
equilibrium, these are the Boltzmann or Fermi functions. Under the influence of a
field, this distribution is modified. It is the purpose of an advanced theory to
determine the modified distribution function. From it, other transport parameters
can be derived.
A formalism first proposed by Liouville (1838, see Ferziger and Kaper 1972) is
too cumbersome for the evaluation of carrier transport. A more useful approach can
be derived from the Liouville equation as a zeroth-order approximation18; this was
suggested by Boltzmann, based on empirical arguments and will be described below.
Conventionally, one uses an accounting procedure for carriers in phase space,
i.e., in a six-dimensional space-and-momentum representation (x, y, z, kx, ky, kz). The
population of electrons in phase space is given by the distribution function f (r, k, t);
it changes with time. A group of electrons within a volume element of phase space
will move and reside in different volume elements as time progresses. Such motion is
described by df/dt. To express the total differential by the local differential, one must
consider the deformation of the distribution function due to the time dependence of
r and k and obtains, using only the first term of a Taylor expansion:

 
df @f @f @f @f
¼ þ k_  þ r_  ¼ ; (83)
dt @t @k @r @t coll

where the index coll indicates the collisions. The first term accounts for the local
change of the distribution in time, the second term for the change in momentum
space, and the third term for the change of the distribution in real space. The sum of
these changes must be equal to the changes of the distribution caused by collisions.
This simplified Liouville equation is called the Boltzmann equation.
In steady state (@f/@t 0), the Boltzmann equation reads

18
The most severe approximation is the linear relation in time, which eliminates memory effects in
the Boltzmann equation (Nag 1980).
34 K. W. Böer and U. W. Pohl

 
@f ðr, k, tÞ @f @f
¼ k_  þ r_  (84)
@t coll @k @r

with the first term determined by the forces acting on free electrons, and k_ given by

e
k_ ¼  F: (85)

Here F is the electric field. The second term is proportional to the spatial gradient of
the carrier distribution and to the group velocity

1 @EðkÞ
r_ ¼ ¼ v: (86)
ℏ @k
This basic Boltzmann equation contains all the dependences necessary for analyzing
carrier transport.19 Some of these dependences, such as the temperature dependence,
are contained implicitly. The important part for the carrier transport is the innocent-
looking left-hand side of the Boltzmann Eq. 84, which contains the contribution of
the more or less inelastic collisions that provide the “friction” for the carrier
transport.
The collision term, also referred to as the collision integral, describes the transi-
tion of an electron from a state Ek, k to a state Ek0 , k0 . This can be expressed as the
difference between electrons scattered from the state k, occupied according to the
Fermi-Dirac distribution function fFD(k), into the state k0 , unoccupied according to
1  f FD ðk0 Þ, minus the reverse process, and integrated over all possible states k0 , into
and from which such scattering is possible:
  ð
@f ðkÞ V
¼ ff FD ðkÞ½1  f FD ðk0 ÞSðk, k0 Þ  f FD ðk0 Þ½1  f FD ðkÞSðk0 ,kÞg dk0 ;
@t coll ð2π Þ3
(87)
where V is the crystal volume and S is the scattering probability


Sðk, k0 Þ ¼ jMðk, k0 Þj δðEk  Ek0  ΔEÞ:
2
(88)

19
Here discussed for electrons, although with a change of the appropriate parameters, it is directly
applicable to holes, polarons, etc. The influence of other fields, such as thermal or magnetic fields, is
neglected here; for such influences, see chapter ▶ “Carriers in Magnetic Fields and Temperature
Gradients”.
Carrier-Transport Equations 35

M(k, k0 ) is the matrix element for the scattering event and ΔE is the fractional
change in electron energy during the partially inelastic scattering. The matrix
elements can be expressed as
ð
Mðk, k0 Þ ¼ ψ q0 , k0 ΔVψ q, k dv (89)
v

where ΔV is the perturbation potential inducing the scattering event, dv is the volume
element, and ψ q, k , ψ q0 , k0 are the wavefunctions before and after scattering. The
perturbation potential depends on the type of scattering event and could be the
deformation potential for scattering on acoustic or optical phonons; see chapter
▶ “Carrier Scattering at Low Electric Fields”:

Ξ c ð@=@rÞ  u acoustic phonons
ΔV ¼ (90)
D0 u optical phonons

with u as the displacement of the lattice atoms and Ξ c or D0 as the appropriate


deformation potentials. Other examples will be given in chapter ▶ “Carrier Scatter-
ing at Low Electric Fields”, where also the scattering potential and matrix elements
for some of the most important scattering centers are tabulated.

4.2 The Boltzmann Equation for Phonons

A Boltzmann-type equation similar to Eqs. 83 and 84 can be set up for the phonon
system, which interacts with the electron system. Since the only driving forces for
the phonon system are those of diffusion due to thermal gradients (neglecting drag
effects discussed in Sect. 4.7), we obtain for steady-state conditions
 
@f ðqÞ @f ðqÞ @ωðqÞ
¼ r_  with r_ ¼ ; (91)
@t coll @r @r

here r_ is the group velocity of phonons, i.e., the sound velocity in the low
q acoustic branch. The gradient of the phonon distribution function f (r, q, T(r),
t) contains the thermal gradient. If undisturbed, the phonon distribution
is described by the Bose-Einstein function fBE(q) – see Eq. 12 of chapter ▶ “Equi-
librium Statistics of Carriers”. The collision term contains all phonon-phonon and
phonon-electron interactions. We will regard the former as less important to the
present discussion. Interaction of optical with acoustic phonons, however, can
become quite important, e.g., for cooling of a heated electron ensemble – see chapter
▶ “Dynamic Processes”. In a fashion similar to that given for the electron collision
term (Eq. 87), we obtain the transition rate by taking the product of the densities of
the occupied and the empty states and the matrix element for each transition,
integrated over all possible transitions for absorption and a similar term for emission
of phonons,
36 K. W. Böer and U. W. Pohl

  ðð
@f ðk, qÞ V
¼ fSðk þ q, kÞ½1 þ f BE ðkÞ f BE ðk þ qÞ g
@t coll ð2π Þ3
 Sðk, k þ qÞ ½1 þ f BE ðk þ qÞ f BE ðkÞ
dk dq : (92)

In equilibrium, the right side vanishes as transitions from k to k + q equal those


from k + q to k. Only when a perturbation is introduced, either from the electron
ensemble interacting with phonons or from a temperature gradient, will the right side
remain finite. The collision term dealing with the interaction of phonons and
electrons can be evaluated after linearization.
In order to obtain numerical values, however, one needs to introduce specific
assumptions about the microscopic collision process between phonons and elec-
trons. The analysis of such collisions will fill the major part of chapter ▶ “Carrier
Scattering at High Electric Fields”.

4.3 The Relaxation-Time Approximation

In order to further discuss carrier transport, we have to solve the Boltzmann


equation; that is, we have to obtain an expression for f (r, k,t). Since the Boltzmann
equation is a nonlinear integrodifferential equation, it cannot be integrated analyti-
cally and requires the use of approximations or of numerical methods. Both will be
mentioned later (Sect. 4.4). However, in order to see some of the important relations,
a simplified approach is introduced first: the relaxation-time approximation.
The balance between gain due to all forces and loss due to collisions of a
perturbation, induced by external forces, produces a steady state with a deformed
electron distribution. When such forces are suddenly removed, the distribution
rapidly returns to its unperturbed state according to
 
@f @f
¼ : (93)
@t @t coll

Assuming that this collision term is linear in the deviation from the unperturbed
distribution f0, we obtain
 
@f f  f0
¼ : (94)
@t coll τm

Equation 93 can then be integrated. It yields an exponential return from the steady-
state, perturbed function f = f0 + δ f (with a time-independent δ f ) to the undisturbed
distribution in equilibrium f0 with the momentum relaxation time τm as the charac-
teristic time constant (see Sect. 4.6.1):
Carrier-Transport Equations 37

 
t
f ðtÞ  f 0 ¼ δf exp  : (95)
τm

In this linearized form, the deformed distribution function will be used first. In the
following sections, an example with zero magnetic field and vanishing gradients in
n and T is discussed. Here the second term of Eq. 84 vanishes.
In a homogeneous semiconductor with a force produced by a constant electric field
F, we obtain from Eqs. 94 and 84 for a small perturbation of the distribution function

eF @f δf
¼ (96)
ℏ @k τm
or, with ℏk ¼ mn v, hence @=@k ¼ ðℏ=mn Þ @=@v, we obtain

e @f
τm F ¼ δ f: (97)
mn @v
This shows that the change in the distribution function is proportional to the drift
velocity vd = (e/mn) τmF. With f0 given by the Boltzmann distribution
 m n v2

f 0 / exp  2
; (98)
kT

we obtain for a small perturbation

@f @f mn v
ffi 0¼ f ; (99)
@v @v kT 0
and with Eq. 97, we have the following equation as the final result for the deformed
Boltzmann distribution due to an external field, in the relaxation time approximation:

 e 
f ¼ f0 1  τm F  v : (100)
kT

An illustration of the relaxation process within this approximation is given in


Fig. 13. A constant and small homogeneous electric Field F shifts the Fermi sphere
in k space a small amount δk along the field direction. After switching off the field,
the return to equilibrium proceeds by inelastic scattering processes from occupied
states of the displaced sphere to unoccupied states of the undisturbed distribution in
equilibrium. Therefore, only electrons in states with jkj > kF scattered to states with
jkj < kF can reestablish equilibrium, with kF being the radius of the Fermi sphere
illustrated in Fig. 13. These are all electrons with energy close to the Fermi energy
EF, which have states near the surface of the Fermi sphere and hence a speed near
vF ¼ ℏkF =mn .
38 K. W. Böer and U. W. Pohl

4.4 Carrier Scattering and Energy Relaxation

For a homogeneous semiconductor, @f =@r 0 holds, and we obtain from Eqs. 83 and 87

@f ðkÞ e F @f ðkÞ
¼ 
@t ℏ @k
ð
V
 3 ff ðkÞ½1  f ðk0 Þ Sðk, k0 Þ  f ðk0 Þ½1  f ðkÞ Sðk0 , kÞg dk0 ; (101)

with the electric field F producing a deformation from the equilibrium distribution
and the collision integral, i.e., the second term in Eq. 99, counteracting this defor-
mation. For an analysis, see Haug (1972).
In equilibrium, the solution of the Boltzmann equation is the Boltzmann or Fermi
function f0(k). With an applied electric field F, the distribution is shifted by the drift
velocity (see Fig. 13) and is slightly deformed. If the effective mass mn is isotropic,
the scattering probability S (Eq. 88) generally depends only on the magnitude
jk  k0 j and hence on the angle included by the vectors, but not on the individual
orientations of k and k0 . The distribution function is then conveniently expressed by
a series development using Legendre polynomials (Pn):

X
1
f ðkÞ ¼ f 0 ðkÞ þ f n ðkÞ Pn ð cos θÞ; (102)
n¼1

where θ is the angle between F and k. After introducing f (k) into Eq. 101, we obtain
for the steady-state (@f =@t 0) a set of n equations to determine f (k).

ky

kx

δkx

Fig. 13 At an applied electric field, the Fermi sphere of occupied states in k space (green-filled
circles) is shifted by an amount δkx, yielding the sphere indicated by the red circle; magenta filling
indicates states occupied due to the applied field. Relaxation to equilibrium occurs by inelastic
scattering into unoccupied states (red arrow) of the equilibrium sphere (gray circle)
Carrier-Transport Equations 39

For small fields, only the first two terms of the development of Eq. 102 are taken
and yield

f ðkÞ ¼ f 0 ðkÞ þ cos θ f 1 ðkÞ: (103)

The perturbation term f1(k) of the distribution function is often expressed in terms of
a function ϕ(E)

eℏ @f 0 ðEÞ
f 1 ðkÞ ¼ Fk ϕðEÞ; (104)
mn @E
with E = E(k). This permits a simplified expression for the collision integral (Nag
1980)
  ð
@f ðkÞ eℏ V
@t ¼  Fk f ð EÞ ½ 1  f 0 ð E0 Þ 
coll mn k T 8π 3 0
  (105)
k0 cos θk

ϕ ð EÞ  ϕðE0 Þ Sðk, k0 Þ dk0 ;
k

where θk is the angle between k and k0 and E0 is the energy corresponding to the
wavevector k0 .
Any further simplification of the collision integral requires assumptions of the
specific scattering event, which will be listed in Sect. 1 of chapter ▶ “Carrier
Scattering at Low Electric Fields” and dealt with sequentially in the following
sections. However, some general remarks here will assist in categorizing the differ-
ent scattering types.

Elastic Scattering Elastic scattering keeps the electron energy during the scattering
event unchanged: E0 ¼ E. This simplifies Eq. 105 to
 
@f ðkÞ eℏ
¼ Fk
@t coll mn k T
ð
V

f 0 ð EÞ ½ 1  f 0 ð EÞ  3 ð1  cos θk Þ Sðk, k0 Þ dk0 : (106)

Elastic scattering events are:

1. All acoustic phonon scattering events, such as deformation-potential scattering


and piezoelectric scattering
2. All defect scattering events, such as scattering at neutral impurities, ionized
impurities, and larger-defect scattering
3. Alloy scattering
40 K. W. Böer and U. W. Pohl

Inelastic Scattering Inelastic scattering does not permit further simplification of


Eq. 105. Here E0 6¼ E , and in each case the collision integral must be evaluated
separately. Such inelastic scattering events are

1. Optical phonon scattering, such as nonpolar and polar optical scattering


2. Intervalley scattering

The total scattering term is given as the sum over the different scattering types:
  X @f 
@f
¼ : (107)
@t coll i
@t coll, i

The Carrier Current The current of carriers can be obtained from the
deformed Boltzmann distribution Eq. 100 by summation over all nel carriers
and velocities v,
X
nel X Xnel X
v F  v f 0 τm
j¼e e v δf ¼ e2 : (108)
i¼1 v i¼1 v
kT

Assuming a spherical equi-energy surface for E(k), the summation over v F  v can be

carried out,20 using for v v the averages vx vy ¼ vy vz ¼ hvz vx i ¼ 0 and v2x
D E
¼ v2y ¼ v2z ¼ v2 =3; this yields

e2 X 2
j¼ v τm f 0 F: (109)
3k T v

Considering that

X
nel X
n¼ f 0; (110)
i¼1 v

we obtain
X
v2 τ m f 0
eF v e 2
j ¼ en X ¼ en v τm F; (111)
3k T f0 3k T
v

20
The quantities τm and f0 in Eq. 108 are functions of E only and hence do not change over a
constant-energy surface.
Carrier-Transport Equations 41

which gives the electron mobility as

e 2
μn ¼ v τm : (112)
3k T

With hEi ¼ ð3=2Þ k T ¼ mn v2 =2; we can replace 3 kT by mnhv2 i, yielding


e v2 τ m e hE τ m i
μn ¼ ¼ : (113)
mn hv2 i mn hEi

This result replaces the average time between scattering events obtained from the
Drude theory with the energy-weighted average of the relaxation time τm. Dropping
the requirement of spherical equi-energy surfaces, the end result (Eq. 113) remains
the same, except that mn is replaced by the anisotropic mobility effective mass; for
more detail, see Conwell (1982).

4.5 The Mobility Effective Mass

Following external forces, the carriers are accelerated proportionally to their effec-
tive masses (Eq. 23). Their anisotropy is taken into consideration by introducing a
mobility effective mass. For a three-axes ellipsoid, this effective mass is given by the
inverse average of the effective masses along the main axes:
 
1 1 1 1 1
¼ þ þ : (114)
mn, μ 3 m1 m2 m3

In general, a mobility tensor is introduced for each of the νd satellite valley E(k)
ellipsoids, identified by the index i:
0 1
@ 2 EðiÞ @ 2 EðiÞ @ 2 EðiÞ
B C
B @k2x @kx @ky @kx @kz C
B 2 ðiÞ C
e τm B @ E @ 2 EðiÞ @ 2 EðiÞ C
μðiÞ ¼ 2 B B
C: (115)
ℏ B @ky @kx @k2y @ky @kz CC
B 2 ðiÞ C
@@ E @ 2 EðiÞ @ 2 EðiÞ A
@kz @kx @kz @ky @k2z

In Si, there are three pairs of ellipsoids with different h1 0 0i orientations;


see Fig. 10a of chapter ▶ “Bands and Bandgaps in Solids”.” All of these
ellipsoids have their E(k) minima at equal energies and are therefore equally
ðiÞ
populated at vanishing external forces with N 0 ¼ 16N 0 . Their effect on the total
mobility is obtained by adding its components, which results in an isotropic
mobility tensor
42 K. W. Böer and U. W. Pohl

0   1
1 2 1
þ 0 0
B 3 mn, t mn, l C
B   C
e τm B 1 2 1 C
μtot ¼ 2 B 0 þ 0 C; (116)
ℏ B 3 mn, t mn, l C
B  C
@ 1 2 1 A
0 0 þ
3 m n, t m n, l

with indices l and t indicating transversal and longitudinal components with


respect to the h1 0 0i axes. Consequently, the electron mobility effective mass
for Si is given by

 
1 1 2 1
¼ þ : (117)
mn, μ 3 mn, t mn, l

The hole mobility effective mass can be derived in a similar fashion. Assuming
spherical E(k) surfaces around k = 0 (for warped bands, see below) and disregarding
the deeper spin-orbit band, one obtains

 
1 1 1 1
¼ þ ; (118)
mp, μ 2 mp, lh mp, hh

here, mp,lh and mp,hh represent the light and heavy hole masses in the corresponding
bands.

Hole-Mobility Mass in Warped Bands The valence bands are significantly


warped; see Sects. 1.2.2 and 1.2.3 of chapter ▶ “Bands and Bandgaps in Solids”.
In cubic semiconductors, the E(k) surfaces can be represented by an empirical
expression (Eq. 15 of chapter ▶ “Bands and Bandgaps in Solids”); using the
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
abbreviations B0 ¼ B2 þ 16C2 and Γ ¼ C2 =f2B0
ðA  B0 Þg in this equation,
we obtain, after taking the second derivative of E(k) for the effective masses, a useful
approximation

m0

m p,  ¼ 0 1 þ 0:333 Γ þ 0:0106 Γ þ . . .
2
(119)
AB
for the mobility effective mass. The variable m represents the light or heavy
hole mass if the upper or lower sign, respectively, is used in Eq. 119. The values
of the constants A, B, and C can be obtained from the Luttinger parameters
(Eq. 16 of chapter ▶ “Bands and Bandgaps in Solids”), which are given in
Table 5 of chapter ▶ “Bands and Bandgaps in Solids” for a number of typical
semiconductors.
Carrier-Transport Equations 43

4.6 Momentum and Energy Relaxation

With each scattering event, momentum is exchanged; the carrier changes the direc-
tion of its path. In addition, more or less energy is exchanged, with the carrier losing
or gaining energy from the scattering center. There are different rates for momentum
and energy relaxation, and the relaxation times depend on the specific scattering
mechanism.

4.6.1 The Average Momentum Relaxation Time


The average momentum relaxation time, which was already introduced in Eq. 95, is
defined as
ð ð
τm ðvÞ vx ð@f =@vx Þ d3 v τm v2x f 0 d 3 v
mn
hτ m i ¼ ð ¼ ð : (120)
3 kT
f0d v f 0 d3 v

It can be obtained from the net increment of the electron momentum, which is
proportional to the average drift velocity,
ð
v x f ð v Þ gð v Þ d 3 v
vx ¼ ð ; (121)
f ðvÞ gðvÞ d 3 v

where g(v) is the density of states. Assuming only small changes from the thermal
distribution, g(v) can be expressed as the effective density of states at the edge of the
band (Eq. 18 of chapter ▶ “Equilibrium Statistics of Carriers”) and cancels out in
Eq. 121. This yields
ð1
v4 τm f 0 ðvÞ dv
e Fx 0 e Fx 2
vx ¼  ð1 ¼ v τm ðvÞ : (122)
3kT 3kT
v2 f 0 ðvÞ dv
0

Using the equipartition law for a Boltzmann gas of electrons m2 v2 ¼ 32k T, we obtain
for the average drift velocity

e Fx v 2 τ m ð v Þ e
vx ¼  ¼  h τ m i Fx : (123)
3kT hv2 i mn

This result is closely related to the Drude equation, however, having replaced τ in
Eq. 25 with the average momentum relaxation time
44 K. W. Böer and U. W. Pohl

2
v τ m ðvÞ
hτm i ¼ : (124)
hv2 i

For an evaluation of Eq. 124, we need the distribution function f0 and the actual
scattering mechanism to determine τm(v) – see chapter ▶ “Carrier Scattering at Low
Electric Fields” and Seeger (1973).
After a collision, the electron path changes by an angle θ, and the fractional
change of angle per collision is on the average h1  cos θi . The momentum
relaxation time is the time after which the electron path is totally randomized, i.e.,
its “memory” is lost; hence,

τsc
τm ¼ ; (125)
h1  cos θi

where τsc is the average time between two collisions. Scattering with θ  90 is
memory erasing (Sect. 3.1 of chapter ▶ “Carrier Scattering at Low Electric
Fields”). Only the collisions in which all angles θ are equally probable result
in hcos θi = 0 and, therefore, yield τm = τsc. For small-angle scattering events,
one needs several scatterings before the momentum is relaxed, leading to τm >
τsc.

The Mean Free Path of Carriers Between collisions, the carrier traverses one free
path. The mean free path is obtained by averaging
2
v λðvÞ
λ¼ : (126)
hv 2 i

λ is related to the momentum relaxation time:

2 rffiffiffiffiffi
v 3π
λ ¼ hτ m i ¼ hτm ivrms : (127)
hvi 8

4.6.2 The Average Energy Relaxation Time


The energy loss or gain due to scattering of electrons with phonons is given by
ð
dE V n    o
¼ 3 ℏωq Sðk, k0 Þ emi, q  ℏωq Sðk0 , kÞ abs, q dk0 : (128)
dt 8π
The subscript emi,q stands for emission and abs,q for absorption of a phonon. When
multiplying the Boltzmann equation with E and integrating over k, we obtain
(Seeger 1973)
Carrier-Transport Equations 45

d hEi h Ei  E L
¼ e F h vi  ; (129)
dt τe
with EL the equilibrium energy at lattice temperature,
ð ð
E f dk v f dk
h Ei ¼ ð , and hvi ¼ ð : (130)
f dk f dk

The energy relaxation time is then obtained from Eq. 127 after switching off the
field, yielding

hEi  EL
τe ¼ ; (131)
d hE i
dt
this means, τe is given by the ratio of the average surplus energy to the rate of energy
loss due to scattering and is a function of E; the energy-loss rate is not a simple
exponential function. It shows a maximum when the electron energy equals the
optical phonon energy.
The rate of momentum or energy loss depends on the actual scattering mecha-
nism. From gas-kinetic arguments, one obtains for collisions between an electron
and a lattice defect of mass M an energy-exchange rate of
 
τe mn
¼ : (132)
τsc ion M

The energy loss is negligible in one scattering event if the scattering center is a defect
atom, since M  mn.
The fraction of energy lost by an electron in a collision with acoustic phonons can
also be obtained from an effective-mass ratio. Using the equivalent phonon mass

kT
mphonon ¼ ; (133)
v2s

where vs is the sound velocity, we obtain


 
τe mn mn v2s 3v2
¼ ¼ ¼ 2s ; (134)
τsc ac phonon mphonon kT vrms

which is on the order of 103. In other words, only 0.1% of the electron energy can
be lost to an acoustic phonon during any one-scattering event.
In contrast, the ratio of energy relaxation time to scattering time for optical
phonons is
46 K. W. Böer and U. W. Pohl

 
τe ℏω0
¼ ; (135)
τsc opt phonon kT

which is on the order of 1 at room temperature. This means that many scattering
events usually pass before the accumulated energy obtained from the field can be
dissipated to the lattice by emitting one optical phonon, while the momentum is
relaxed after one or only a few collisions. This modifies the rather crude model
given in Sect. 2.1 by introducing the momentum relaxation time for evaluating the
mobility and the energy relaxation time for Joule’s heating.21 A more detailed
discussion is given in chapters ▶ “Carrier Scattering at Low Electric Fields” and
▶ “Carrier Scattering at High Electric Fields”, when an analysis of the different
scattering mechanisms and a better estimate of the magnitude of energy obtained
from the field is given.

4.7 Phonon and Electron Drag

Interacting electrons and phonons exchange energy and momentum. A drift motion
superimposed on the random motion of one ensemble transfers part of the net
momentum to the other ensemble during scattering. This means that electrons
drifting in an external field tend to push phonons in the same direction, which causes
a slight temperature gradient in the field direction, superimposed on the homoge-
neous Joule’s heating. This process is called electron drag (Hubner and Shockley
1960).
Similarly, a temperature gradient tends to push electrons from the warm to the
cold end of a semiconductor. This is known as phonon drag. The drag effect can be
quite large, e.g., up to a factor of 6 compared to simple thermopower in p-type Ge at
20 K, as shown by Herring (1954).
When phonons propagate as acoustic waves, ac-electric fields can be induced; or,
vice versa, when sufficiently high electric fields are applied, coherent phonon waves
can be generated when the drift velocity of electrons surpasses the (sound) velocity
of the phonon waves (McFee 1966). These acousto-electric effects have technical
application for creating current oscillators (Bray 1969).

5 Summary

Electrons and holes are carriers of the current in semiconductors. Their effective
mass is modified by polarizing the lattice, but this influence, expressed as a change to
polarons, is negligible for most semiconductors. It is contained in the effective mass

21
In two-dimensional structures, significant momentum relaxation is found due to near-surface
acoustic phonon scattering; see Pipa et al. (1999).
Carrier-Transport Equations 47

obtained by cyclotron resonance and is commonly listed as the effective mass of the
carriers, i.e., of electrons or holes. Only for small polarons, which occur in semi-
conductors with narrow bands and large lattice polarization, the effective mass is
significantly affected and self-trapping occurs.
The current through a semiconductor is composed of a drift and a diffusion
current of electrons and holes. In homogeneous semiconductors, only one of these
four components is dominant. The drift current is determined by the electric field,
which acts as a slight perturbation of an essentially random walk of carriers, except
for very high fields. The additional energy obtained by carrier acceleration from the
field is given to the lattice by inelastic scattering, causing Joule’s heating. The
diffusion current is proportional to the carrier gradient up to a maximum diffusion
current, which is limited by the thermal velocity of carriers. Proportionality factor of
both drift and diffusion currents is the carrier mobility, which is proportional to a
relaxation time and inversely proportional to an effective mass tensor, the mobility
effective mass.
Drift and total currents are proportional to negative potential gradients: the first
one being the electrostatic potential and the second the electrochemical potential.
The proportionality factor of both is the conductivity. In spatially inhomogeneous
semiconductors, both an external field, impressed by an applied bias, and a built-in
field, due to space-charge regions, exist within the semiconductor. The external field
causes carrier heating by shifting and deforming the carrier distribution from a
Boltzmann distribution to a distorted distribution with more carriers at higher
energies within the band. In contrast, the built-in field leaves the Boltzmann distri-
bution of carriers unchanged; the carrier gas remains unheated at exactly the same
temperature as the lattice at every volume element of the crystal, except for statistical
fluctuations. A consequence of the difference between external and built-in fields is
the difference in determining the field dependence of the mobility, which requires an
averaging over carriers with different energies within the band. For a built-in field,
the averaging follows a Boltzmann distribution; in an external field, there are more
electrons at higher energies, and the distribution is distorted accordingly. This can
have significant impact for the evaluation of device performances when high fields
are considered.
The Boltzmann equation permits a sophisticated analysis of the carrier transport,
including changes in the carrier distribution. Such change in the distribution sub-
stantially influences the averaging, which is necessary to arrive at well-defined
values for a number of transport parameters – most importantly the relaxation
times. Significant differences can be defined between the time between two scatter-
ing events τsc, the momentum relaxation time τm, and the energy relaxation time τe.
Although the Boltzmann equation cannot be integrated in closed form except for a
few special cases, the deformed distribution function can be approximated for small
applied fields. It provides the basis for the investigation of various scattering
processes, which can be divided into essentially elastic processes with mainly
momentum exchange and, for carriers with sufficient accumulated energy, into
inelastic scattering with energy relaxation. The latter becomes more prevalent at
elevated temperatures and higher electric fields.
48 K. W. Böer and U. W. Pohl

References
Appel J (1968) Polarons. In: Seitz F, Turnbull D, Ehrenreich H (eds) Solid state physics, vol 21.
Academic Press, New York, pp 193–391
Bajaj KK (1968) Polaron in a magnetic field. Phys Rev 170:694
Böer KW (1985a) High-field carrier transport in inhomogeneous semiconductors. Ann Phys
497:371
Böer KW (1985b) Current-voltage characteristics of diodes with and without light. Phys Status
Solidi A 87:719
Bogoliubov NN, Bogoliubov NN Jr (1986) Aspects of polaron theory. World Scientific Publishing,
Singapore
Böttger H, Bryksin VV (1985) Hopping conduction in solids. Phys Status Solidi B 96:219
Bray R (1969) A Perspective on acoustoelectric instabilities. IBM J Res Dev 13:487
Brazovskii S, Kirova N, Yu ZG, Bishop AR, Saxena A (1998) Stability of bipolarons in conjugated
polymers. Opt Mater 9:502
Chakraverty BK, Schlenker C (1976) On the existence of bipolarons in Ti4O7. J Physique (Paris)
Colloq 37:C4–C353
Chandrasekhar S (1943) Stochastic problems in physics and astronomy. Rev Mod Phys 15:1
Chattopadhyay D, Rakshit PC, Kabasi A (1989) Diffusion of one- and two-dimensional hot
electrons in semiconductor quantum-well structures at low temperatures. Superlattice Micro-
struct 6:399
Christov SG (1982) Adiabatic polaron theory of electron hopping in crystals: a reaction-rate
approach. Phys Rev B 26:6918
Cohen MH, Economou EN, Soukoulis CM (1983) Electron-phonon interactions near the mobility
edge in disordered semiconductors. J Non-Cryst Solids 59/60:15
Comas F, Mora-Ramos ME (1989) Polaron effect in single semiconductor heterostructures. Physica
B 159:413
Conwell EM (1967) High-field transport in semiconductors. Academic Press, New York
Conwell EM (1982) The Boltzmann equation. In: Paul W, Moss TS (eds) Handbook of semi-
conductors vol 1: band theory and transport properties. North Holland Publishing, Amsterdam,
pp 513–561
Devreese JT (1984) Some recent developments on the theory of polarons. In: Devreese JT, Peeters
FM (eds) Polarons and excitons in polar semiconductors and ionic crystals. Plenum Press,
New York, pp 165–183
Drude P (1900) Zur Elektronentheorie der Metalle. Ann Phys 1:566. (On the electron theory of
metals, in German)
Dushman S (1930) Thermionic emission. Rev Mod Phys 2:381
Emin D (1973) On the existence of free and self-trapped carriers in insulators: an abrupt
temperature-dependent conductivity transition. Adv Phys 22:57
Esaki L, Tsu R (1970) Superlattice and negative differential conductivity in semiconductors. IBM J
Res Dev 14:61
Evrard R (1984) Polarons. In: Di Bartolo B (ed) Collective excitations in solids. Plenum Press,
New York, pp 501–522
Ferziger JD, Kaper HG (1972) Mathematical theory of transport processes in gasses. North Holland
Publishing, Amsterdam
Feynman RP (1955) Slow electrons in a polar crystal. Phys Rev 97:660
Frenkel JI (1936) On the absorption of light and the trapping of electrons and positive holes in
crystalline dielectrics. Sov Phys J 9:158
Fröhlich H, Pelzer H, Zienau S (1950) Properties of slow electrons in polar materials. Philos Mag
(Ser 7) 41:221
Haug A (1972) Theoretical solid state physics. Pergamon Press, Oxford
Hayes W, Stoneham AM (1984) Defects and defect processes in nonmetallic solids. Wiley,
New York
Carrier-Transport Equations 49

Herring C (1954) Theory of the thermoelectric power of semiconductors. Phys Rev 96:1163
Holstein T (1959) Studies of polaron motion Part 1: the molecular-crystal model. Ann Phys 8:325
Horio K, Okada T, Nakatani A (1999) Energy transport simulation for graded HBT’s: importance of
setting adequate values for transport parameters. IEEE Trans Electron Devices 46:641
Hubner K, Shockley W (1960) Transmitted phonon drag measurements in silicon. Phys Rev Lett
4:504
Kan EC, Ravaioli U, Chen D (1991) Multidimensional augmented current equation including
velocity overshoot. IEEE Electron Device Lett 12:419
Kartheuser E, Devreese JT, Evrard R (1979) Polaron mobility at low temperature: a self-consistent
equation-of-motion approach. Phys Rev B 19:546
Landau LD (1933) Electron motion in crystal lattices. Sov Phys J 3:664
Landsberg PT (1952) On the diffusion theory of rectification. Proc R Soc Lond A Math Phys Sci
213:226
Landsberg PT, Cheng HC (1985) Activity coefficient and Einstein relation for different densities of
states in semiconductors. Phys Rev B 32:8021
Lee TD, Low F, Pines D (1953) The motion of slow electrons in a polar crystal. Phys Rev 90:297
Leo K (1996) Optical investigations of Bloch oscillations in semiconductor superlattices. Phys
Scripta T68:78
Liouville J (1838) Note sur la théorie de la variation des constantes arbitraires. J Math Pures Appl
3:349. (Note on the theory of the variation of arbritrary constants, in French)
Lorentz HA (1909) The theory of electrons. Teubner Verlag, Leipzig
McFee JH (1966) Transmission and amplification of acoustic waves. In: Mason W (ed) Physical
acoustics, vol 4A. Academic Press, New York, pp 1–44
Mikhnenko OV, Blom PWM, Nguyen T-Q (2015) Exciton diffusion in organic semiconductors.
Energy Environ Sci 8:1867
Muljarov EA, Tikhodeev SG (1997) Self-trapped excitons in semiconductor quantum wires inside a
polar dielectric matrix. Phys Status Solidi A 164:393
Nag BR (1980) Electron transport in compound semiconductors. Springer, Berlin
Najafi E, Scarborough TD, Tang J, Zewail A (2015) Four-dimensional imaging of carrier interface
dynamics in p-n junctions. Science 347:164
Peeters FM, Devreese JT (1984) Theory of polaron mobility. In: Ehrenreich H, Turnbull D (eds)
Solid state physics, vol 38. Academic Press, Orlando, pp 81–133
Pekar SI (1954) Untersuchungen über die Elektronentheorie der Kristalle. Akademie Verlag, Berlin.
(Investigations on the electron theory of crystals, in German)
Pipa VI, Vagidov NZ, Mitin VV, Stroscio M (1999) Momentum relaxation of 2D electron gas due to
near-surface acoustic phonon scattering. Physica B 270:280
Seeger K (1973) Semiconductor physics. Springer, Wien/New York
Shah J (1994) Bloch oscillations in semiconductor superlattices. Proc SPIE 2145:144
Simoen E, Claeys C, Czerwinski A, Katcki J (1998) Accurate extraction of the diffusion current in
silicon p-n junction diodes. Appl Phys Lett 72:1054
Sommerfeld A (1928) Zur Elektronentheorie der Metalle auf Grund der Fermischen Statistik. Z
Phys 47:1 (On the electron theory due to Fermi’s statistics, in German)
Stratton R (1969) Carrier heating or cooling in a strong built-in electric field. J Appl Phys 40:4582
Tjablikov SV (1952) Zh Eksp Teor Fiz 23:381
Toyozawa Y (1981) Charge transfer instability with structural change. I. Two-sites two-electrons
system. J Phys Soc Jpn 50:1861
Velasco VR, García-Moliner F (1997) Polar optical modes in semiconductor nanostructures. Surf
Sci Rep 28:123
Waschke C, Leisching P, Haring Bolivar P, Schwedler R, Brüggemann F, Roskos HG, Leo K,
Kurz H, Köhler K (1994) Detection of Bloch oscillations in a semiconductor superlattice by
time-resolved terahertz spectroscopy and degenerate four-wave mixing. Solid State Electron
37:1321

You might also like